paper_id
stringlengths 9
16
| version
stringclasses 26
values | yymm
stringclasses 311
values | created
timestamp[s] | title
stringlengths 6
335
| secondary_subfield
sequencelengths 1
8
| abstract
stringlengths 25
3.93k
| primary_subfield
stringclasses 124
values | field
stringclasses 20
values | fulltext
stringlengths 0
2.84M
|
---|---|---|---|---|---|---|---|---|---|
1805.07415 | 1 | 1805 | 2018-05-18T19:44:56 | Effects of dissociation/recombination on the day-night temperature contrasts of ultra-hot Jupiters | [
"astro-ph.EP",
"physics.ao-ph"
] | Secondary eclipse observations of ultra-hot Jupiters have found evidence that hydrogen is dissociated on their daysides. Additionally, full-phase light curve observations of ultra-hot Jupiters show a smaller day-night emitted flux contrast than that expected from previous theory. Recently, it was proposed by Bell & Cowan (2018) that the heat intake to dissociate hydrogen and heat release due to recombination of dissociated hydrogen can affect the atmospheric circulation of ultra-hot Jupiters. In this work, we add cooling/heating due to dissociation/recombination into the analytic theory of Komacek & Showman (2016) and Zhang & Showman (2017) for the dayside-nightside temperature contrasts of hot Jupiters. We find that at high values of incident stellar flux, the day-night temperature contrast of ultra-hot Jupiters may decrease with increasing incident stellar flux due to dissociation/recombination, the opposite of that expected without including the effects of dissociation/recombination. We propose that a combination of a greater number of full-phase light curve observations of ultra-hot Jupiters and future General Circulation Models that include the effects of dissociation/recombination could determine in detail how the atmospheric circulation of ultra-hot Jupiters differs from that of cooler planets. | astro-ph.EP | astro-ph | Draft version May 22, 2018
Typeset using LATEX RNAAS style in AASTeX62
8
1
0
2
y
a
M
8
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
1
4
7
0
.
5
0
8
1
:
v
i
X
r
a
Effects of dissociation/recombination on the day-night temperature contrasts of
ultra-hot Jupiters
Thaddeus D. Komacek1 and Xianyu Tan1
1Lunar and Planetary Laboratory, University of Arizona, Tucson, AZ, 85721 [email protected]
Keywords: methods: analytical - planets and satellites: atmospheres
MAIN BODY
Recent observations of ultra-hot Jupiters have revealed new physical processes at play that do not occur in the
atmospheres of cooler planets. Secondary eclipse observations (Arcangeli et al. 2018; Kreidberg et al. 2018; Mansfield
et al. 2018) have found spectra that appear featureless throughout the 1.1 − 1.7µm region, likely caused by the
continuum opacity due to dissociated hydrogen (Bell et al. 2017; Lothringer et al. 2018; Parmentier et al. 2018).
Recently, it was shown by Bell & Cowan (2018) that when atomic hydrogen is transported from the hot dayside to
the relatively cold nightside of these planets, recombination of atomic H into H2 releases a significant amount of heat
that can warm up the nightside of the planet. Interestingly, recent phase curve observations of the ultra-hot Jupiters
WASP-103b (Kreidberg et al. 2018) and WASP-33b (Zhang et al. 2018) showed relatively small amplitudes, indicative
of processes reducing their day-night temperature contrast.
In this work, we predict the day-night temperature differences of ultra-hot Jupiters using an analytic theory updated
from that developed in Komacek & Showman (2016); Zhang & Showman (2017); Komacek et al. (2017). This theory
scales the primitive equations of meteorology to solve for characteristic day-night temperature contrasts in hot Jupiter
atmospheres. We include the effects of hydrogen dissociation/recombination, including cooling/heating and the change
in mean molecular weight. Specifically, we add the following energy source to the scaled thermodynamic energy
equation (Equation 24 in Komacek & Showman 2016):
Qrecomb
cp
=
U qbondηH (T )
cp(T )a
,
(1)
where U is the wind speed (solved for consistently using the scaled momentum equation), a is the planetary radius,
qbond = 2.14 × 108 J kg
−1 is the hydrogen dissociation energy (Bell & Cowan 2018), ηH (T ) is the mass mixing ratio
of atomic hydrogen calculated from the molar mixing ratio χH (given by the approximation in Equation 5 of Bell
& Cowan 2018) as ηH = χH /(2 − χH ), and cp(T ) is the specific heat capacity of the combined hydrogen molecular
and atomic gas mixture. Additionally, hydrogen dissociation decreases the day-side mean molecular weight, which
increases the day-night pressure gradient. We include this effect by modifying the horizontal geopotential difference to
take into account the difference between the dayside specific gas constant Rday(T ) = R0(1 + ηH (T )) and the nightside
gas constant Rnight = R0, where R0 is the gas constant of H2.
As in Komacek et al. (2017), we solve for the dayside-nightside temperature difference as a function of its two
incident stellar flux and frictional drag timescale. Figure 1 shows our predictions for the
key control parameters:
day-night brightness temperature difference, A = (Tb,day − Tb,night)/Tb,day, both including and ignoring the effects of
dissociation/recombination. Dissociation and recombination cause the day-night temperature contrast to decrease in
the ultra-hot Jupiter regime, in contrast to the case ignoring this effect that shows the opposite trend. Note that for
even hotter planets than that considered here (e.g., KELT-9b), the nightside may begin to be partially dissociated,
potentially breaking our theoretical assumption that the nightside is constituted of purely molecular hydrogen.
We also show observations of the phase curve amplitude in Figure 1, updated from Komacek et al. (2017) to include
WASP-33b and WASP-103b. These two phase curve observations do not show evidence for an increase in phase curve
amplitude at high temperatures. We find that hydrogen dissociation/recombination provides a mechanism to explain
the relatively small observed phase curve amplitudes of these ultra-hot Jupiters, as predicted by Bell & Cowan (2018).
2
Komacek & Tan
Figure 1. Theoretically predicted dayside-nightside temperature contrast at a pressure of 100 mbars for a planet with the radius,
gravity, and rotation rate of HD 209458b (lines, dotted lines do not include dissociation/recombination) and observations (points,
lines without a point show lower limits), plotted as a function of equilibrium temperature.
This theory provides a basic estimate for how dissociation/recombination of hydrogen affects the day-night tem-
perature contrasts of ultra-hot Jupiters as a function of planetary parameters. However, it does not provide details
about the feedback of dissociation/recombination on dynamics, which is crucial for interpreting observed phase curves.
Dissociation of hydrogen on the dayside should cause a strong cooling effect that could affect dynamics near the sub-
stellar point. Additionally, because the hydrogen recombination timescales in hot Jupiter atmospheres are extremely
short (Parmentier et al. 2018), the majority of recombination might occur in localized regions with large temperature
contrasts (e.g., the limb). As a result, the recombination heating may occur at the limb rather than throughout the
nightside as assumed here. Dynamics could then transport heat from the limb toward the nightside of the planet.
General Circulation Models that include the effects of hydrogen dissociation/recombination can test our theory, and
would provide more detailed comparisons to phase curve observations of ultra-hot Jupiters.
We thank Nick Cowan and Taylor Bell for a careful reading of the manuscript and Josh Lothringer for helpful
discussions.
10001500200025003000Equilibrium temperature0.00.20.40.60.81.0Normalized day-night temperature contrastdrag=drag=104 sNo dissociation3.6m4.5m8m24mWFC3REFERENCES
3
Arcangeli, J. et al. 2018, The Astrophysical Journal
Komacek, T., Showman, A., & Tan, X. 2017, The
Letters, 855, L30
Bell, T.J. et al. 2017, The Astrophysical Journal Letters,
847, L2
Bell, T.J. & Cowan, N.B. 2018, The Astrophysical Journal
Letters, 857, L20
Komacek, T. & Showman, A. 2016, The Astrophysical
Journal, 821, 16
Astrophysical Journal, 835, 198
Kreidberg, L. et al. 2018, arXiv e-prints:1805.00029
Lothringer, J.D., Barman T., & Koskinen T. 2018, arXiv
e-prints:1805.0038
Mansfield, M. et al. 2018, arXiv e-prints:1805.00424
Parmentier, V. et al. 2018, arXiv e-prints:1805.00096
Zhang, M. et al. 2018, The Astronomical Journal, 155, 83
Zhang, X. & Showman, A. 2017, The Astrophysical
Journal, 836, 73
|
1612.06920 | 1 | 1612 | 2016-12-20T23:27:13 | Non-Gravitational Acceleration of the Active Asteroids | [
"astro-ph.EP"
] | Comets can exhibit non-gravitational accelerations caused by recoil forces due to anisotropic mass loss. So might active asteroids. We present an astrometric investigation of 18 active asteroids in search of non-gravitational acceleration. Statistically significant (signal-to-noise ratio (SNR) $> 3$) detections are obtained in three objects: 313P/Gibbs, 324P/La Sagra and (3200) Phaethon. The strongest and most convincing detection ($>$7$\sigma$ in each of three orthogonal components of the acceleration), is for the $\sim$1 km diameter nucleus of 324P/La Sagra. A 4.5$\sigma$ detection of the transverse component of the acceleration of 313P/Gibbs (also $\sim$1 km in diameter) is likely genuine too, as evidenced by the stability of the solution to the rejection or inclusion of specific astrometric datasets. We also find a 3.4$\sigma$ radial-component detection for $\sim$5 km diameter (3200) Phaethon, but this detection is more sensitive to the inclusion of specific datasets, suggesting that it is likely spurious in origin. The other 15 active asteroids in our sample all show non-gravitational accelerations consistent with zero. We explore different physical mechanisms which may give rise to the observed non-gravitational effects, and estimate mass-loss rates from the non-gravitational accelerations. We present a revised momentum-transfer law based on a physically realistic sublimation model for future work on non-gravitational forces, but note that it has little effect on the derived orbital elements. | astro-ph.EP | astro-ph |
Accepted by the Astronomical Journal
Non-Gravitational Acceleration of the Active Asteroids
Man-To Hui (許文韜)1 and David Jewitt1,2
1Department of Earth, Planetary and Space Sciences, UCLA, 595 Charles Young Drive
East, Los Angeles, CA 90095-1567
2Department of Physics and Astronomy, UCLA, 430 Portola Plaza, Box 951547, Los
Angeles, CA 90095-1547
[email protected]
ABSTRACT
Comets can exhibit non-gravitational accelerations caused by recoil forces due
to anisotropic mass loss. So might active asteroids. We present an astrometric in-
vestigation of 18 active asteroids in search of non-gravitational acceleration. Sta-
tistically significant (signal-to-noise ratio (SNR) > 3) detections are obtained in
three objects: 313P/Gibbs, 324P/La Sagra and (3200) Phaethon. The strongest
and most convincing detection (>7σ in each of three orthogonal components of
the acceleration), is for the ∼1 km diameter nucleus of 324P/La Sagra. A 4.5σ
detection of the transverse component of the acceleration of 313P/Gibbs (also
∼1 km in diameter) is likely genuine too, as evidenced by the stability of the so-
lution to the rejection or inclusion of specific astrometric datasets. We also find a
3.4σ radial-component detection for ∼5 km diameter (3200) Phaethon, but this
detection is more sensitive to the inclusion of specific datasets, suggesting that it
is likely spurious in origin. The other 15 active asteroids in our sample all show
non-gravitational accelerations consistent with zero. We explore different physi-
cal mechanisms which may give rise to the observed non-gravitational effects, and
estimate mass-loss rates from the non-gravitational accelerations. We present a
revised momentum-transfer law based on a physically realistic sublimation model
for future work on non-gravitational forces, but note that it has little effect on
the derived orbital elements.
Subject headings:
asteroids: general
comets: general -- methods: data analysis -- minor planets,
-- 2 --
1.
INTRODUCTION
Active asteroids have the dynamical characteristics of asteroids but exhibit transient
mass loss, resulting in the production of comet-like appearance (Hsieh and Jewitt 2006).
A working definition is that they are bodies which present evidence of mass loss, have
semimajor axes, a, smaller than Jupiter's semimajor axis, and have Tisserand parameter
with respect to Jupiter, TJ ≥ 3.08. There are currently ∼20 known active asteroids. A
number of mechanisms drive the mass loss, including the likely sublimation of exposed ice,
asteroid-asteroid impact, and rotational disruption probably driven by radiation torques
(Jewitt 2012; Jewitt et al. 2015).
The dynamics of active asteroids are of particular interest. Numerical simulations have
been conducted to study the dynamical stability of some of these objects (c.f. Jewitt et
al. 2015 and citations therein). Recent work by Hsieh & Haghighipour (2016) investigated
orbital evolution of test particles dynamically close to the TJ ≃ 3 boundary between asteroids
and comets. They found that, due to gravitational interactions with terrestrial planets and
temporary trapping by mean-motion resonances with Jupiter, the fraction of the Jupiter-
family comets fortuitously evolved into main-belt like orbits on Myr timescales could be as
large as ∼0.1 -- 1%. However, most such main-belt captures would be transient, and long-term
stable orbits with both small eccentricities and inclinations should be much more rare.
Non-gravitational accelerations, if present, might significantly influence the dynamics
of small bodies. Fern´andez et al. (2002) and Levison et al. (2006) found that capture into
comet 2P/Encke's orbit is possible when assisted by plausible non-gravitational forces from
outgassed material, but takes much longer than the expected outgassing lifetimes of comets.
They suggested that 2P/Encke might have completed this capture while spending most of
its time in a dormant state. Forces due to photon momentum (the Yarkovsky effect (e.g.,
Chesley et al. 2003; Vokrouhlick´y et al. 2008; Chesley et al. 2012; Nugent et al. 2012;
Farnocchia et al. 2014) and radiation pressure) are expected to be tiny compared to forces
resulting from protracted anisotropic mass loss but have been detected in small asteroids.
To date, the only independently reported measurement of non-gravitational acceleration
due to outgassing in an active asteroid is a 3σ detection for 133P/(7968) Elst-Pizarro (Chesley
et al. 2010a). In order to develop a better understanding of the active asteroids, we attempt
to measure their non-gravitational accelerations.
-- 3 --
2. DATA ANALYSIS AND METHOD
Marsden et al. (1973) developed a standard orbit determination technique with non-
gravitational effects. The non-gravitational acceleration of a small body, in terms of its
radial (i.e., in the antisolar direction), transverse, and normal components AR, AT, and AN,
is related to three non-gravitational parameters Aj (j = 1, 2, 3), which are expressed in the
same right-handed Cartesian orthogonal coordinates system by
AR
AT
AN
=
A1
A2
A3
· g (r) ,
(1)
where g (r) is the dimensionless standard momentum-transfer law at heliocentric distance,
r, in AU. Marsden et al. (1973) defined g(r) as:
g (r) = α(cid:18) r
r0(cid:19)−m(cid:20)1 +(cid:18) r
r0(cid:19)n(cid:21)−k
,
(2)
in which m = 2.15, n = 5.093, k = 4.6142, the scaling distance r0 = 2.808 AU, and the
normalisation factor α = 0.111262, such that g = 1 at r = 1 AU. Accelerations Aj and
Aj are traditionally expressed in AU day−2. The momentum-transfer law comes from the
assumption by Marsden et al.
(1973) that the non-gravitational acceleration of a small
body is proportional to the rate of sublimation of water-ice on an isothermal nucleus, with
the momentum-transfer law reflecting the proportionality, such that the non-gravitational
parameters Aj are always constant. (Sublimation of other materials such as sodium and
forsterite can be approximated by the same formalism with different parameters (c.f. Sekan-
ina & Kracht 2015), but the sublimation rates of these much less volatile materials are
negligible compared to that of water.) In keeping with previous work, we proceed by assum-
ing that the momentum-transfer law due to isothermal water-ice sublimation gives rise to
the non-gravitational effects of the active asteroids.
We downloaded astrometric observations of all the active asteroids from the Minor
Planet Center (MPC) Database Search1, and then employed Find Orb by B. Gray for orbit
determination. The code uses numerical ephemeris DE431, and includes relativistic effects
due to the gravity of the Sun, and perturbations by the eight major planets. Pluto and the
1http://www.minorplanetcenter.net/db_search
-- 4 --
thirty most massive asteroids2 are also included. Astrometric observations were debiased
and weighted as described in Farnocchia et al. (2014) and Chesley et al. (2010b) before
orbit determination.
We first calculated purely gravitational orbital solutions for each of the active asteroids,
assuming Aj = 0 (j = 1, 2, 3). Weights would be relaxed to be comparable with correspond-
ing ad hoc astrometric residuals. We next rejected astrometric observations whose residuals
were greater than ±3′′.0 from ad hoc osculating solutions, in an iterative manner. For main-
belt objects, such residuals are large compared to systematic errors from the timing or plate
constant solutions. They may result from centroiding errors possibly due to the faintness or
non-stellar appearance of the object, from interference with background sources or adjacent
cosmic rays or from other, unspecified errors. The threshold was chosen to exclude bad out-
liers while keeping as many data points as possible. Next, we included Aj (j = 1, 2, 3) as free
parameters to be obtained from the best fit orbital solutions. The procedures for filtering
outliers and relaxing weights were applied iteratively until convergence was achieved. This
normally took three to five runs, somewhat dependent upon the quality of data. We finally
recorded the converged orbital solutions along with Aj (j = 1, 2, 3).
3. RESULTS
We summarize the resulting non-gravitational parameters of the active asteroids in
Table 1.
Included are statistically confident detections (SNR > 3) of non-gravitational
accelerations for 324P/La Sagra in all the three components, for (3200) Phaethon in the
radial direction, and for 313P/Gibbs in the transverse direction. The other active asteroids
show no statistically significant evidence (SNR ≤ 3) for non-gravitational effects.
Our non-detection of the radial component of non-gravitational acceleration in 133P/(7968)
Elst-Pizarro contradicts a 3σ detection reported by Chesley et al. (2010a). However, if only
observations prior to 2011 are considered, our result becomes similar to that of Chesley et al.
(2010a). Therefore, we conclude that the reported detection is tied to the specific astromet-
ric dataset employed, and cannot be trusted as real. Likewise, active asteroid 259P/Garradd
shows marginal evidence of a radial non-gravitational acceleration with SNR = 2.97 (see
Table 1). However, the result is found to change wildly depending on the particular as-
trometric observations selected. Moreover, the fit to 259P/Garradd relies on the smallest
number of observations (40, compared to hundreds or thousands for other objects in Table
2The masses of the 30 most massive asteroids range from ∼7 × 1018 kg (375 Ursula) to 9 × 1020 kg (1
Ceres). The values are based on the BC-405 asteroid ephemeris by Baer et al. (2011).
1). Therefore, we do not regard it as a significant detection.
-- 5 --
3.1.
313P/Gibbs
Hui & Jewitt (2015) previously discussed the non-gravitational motion of this ∼1 km
diameter object. We did not debias the astrometric observations and simply set equal weights
to all the data. Nevertheless, the result is consistent with the one in the present work in
which we employed more stringent techniques to weight the data. In this sense, the detection
of A2, at 4.5σ confidence (Table 1) is relatively insensitive to the method by which the
astrometric observations are handled. We thus conclude that it is likely a genuine detection
of the transverse non-gravitational acceleration. Admittedly, in order to strengthen this
conclusion, more observations of the object are desirable.
3.2.
324P/La Sagra
324P/La Sagra shows the strongest non-gravitational acceleration of all the active aster-
oids, with detections >7σ in all three components (see Table 1). The solutions are unlikely to
be caused by contamination from undetected systematics in the astrometry because random
exclusions of large subsets of the astrometric data hardly change the result. For example,
discarding all the data from 2015 leads to no change in the significance of the Aj parameters.
Other tests, including arbitrary assignment of equal weights to all the data, have been made,
without materially changing the result. While the detection of non-gravitational accelera-
tion appears to be secure, the solution is nevertheless somewhat puzzling. In particular, the
radial component, A1, is negative (radial non-gravitational acceleration towards the Sun),
which seems physically unrealistic in the context of sublimation from the hot day-side of
the nucleus. This may indicate that the applied momentum-transfer law by Marsden et al.
(1973) is inappropriate to this case, because the mass-loss rate does not vary symmetrically
with heliocentric distance (or, equivalently, perihelion time) as described by Equation (2)
(see Figure 6 in Jewitt et al. (2016)). Another possibility is that it suggests a circumpolar
or high-latitude active source and certain combinations of the spin-axis orientation of its
nucleus (Yeomans et al. 2004).
-- 6 --
3.3.
(3200) Phaethon
Since the discovery in 1983, asteroid (3200) Phaethon had never been observed to show
any signs of activity until 2009, 2012 and 2016 when it brightened by a factor of two around
perihelion detected by the Solar Terrestrial Relations Observatory (STEREO) spacecraft
(Jewitt & Li 2010; Li & Jewitt 2013; Hui & Li 2016). Intriguingly, we have a SNR = 3.4
detection for its radial non-gravitational parameter A1, which is statistically significant.
Tests such as discarding all observations prior to 1990, or applying an equal weight scheme do
affect the SNR slightly, but always leave SNR ∼ 3. However, we can destroy the significance
of the detection by, for instance, discarding all the data from the discovery epoch to the
mid-1990s. Alternatively, if a much stricter cutoff for astrometric residuals is employed (e.g.
. 1′′.5), resulting in removing observations overwhelmingly from the 1980s and early 1990s,
the SNR shrinks to ∼2 and thus A1 becomes insignificant. We therefore take the conservative
position that the radial non-gravitational component is likely spurious. This is supported
by the observation that (3200) Phaethon remains inactive until it is close to the Sun, where
the activity is likely triggered by some process (thermal fracture, desiccation?) other than
the sublimation of water ice (Jewitt & Li 2010).
4. DISCUSSION
4.1. Test of the Procedure
We conducted another test of the algorithms used by the orbit determination code
Find Orb to be sure that the software does not introduce false detections of non-gravitational
motion. For this purpose, we selected a dozen asteroids ∼10 km in diameter and having
apparent magnitudes, orbits and observational histories similar to the majority of the active
asteroids. The 10 km asteroids, being ∼103 times more massive than the mostly ∼1 km
scale active asteroids (Table 3), are unlikely to exhibit any measurable non-gravitational
acceleration and thus serve as tests of the orbital fitting. A list of candidates was generated
by the JPL Small-Body Database Search Engine3. We applied the same procedures and
techniques described in Section 2 to obtain orbital solutions including Aj (j = 1, 2, 3) as free
parameters. The results are summarized in Table 2.
As expected, none of the asteroids show significant (>3σ) non-gravitational parameters.
Some of the active asteroids have fewer observations than have the selected moderate sized
3http://ssd.jpl.nasa.gov/sbdb_query.cgi. Data retrieved on 2016 July 14.
-- 7 --
asteroids. We therefore truncated all the observations prior to 2010 for each of these asteroids
and re-performed orbit determination. Again none shows detections on the non-gravitational
parameters with SNR > 3. This confirms past work done with Find Orb (e.g., Micheli et
al. 2014) independently showing the reliability of the code. The validity of our cutoff set at
SNR = 3 is justified as well.
4.2. Mass-Loss Estimates
The mass-loss rate needed to provide a given non-gravitational acceleration can be
estimated thanks to momentum conservation, using
M (t) = −
M (t) g (r (t))pA2
κ (t) v (t)
1 + A2
2 + A2
3
,
(3)
where M is the mass of the body, v is the outflow speed of the ejecta, and κ is a dimensionless
factor which accounts for the collimation efficiency. The latter lies in the range 0 ≤ κ ≤
1, with κ = 0 for isotropic ejection and κ = 1 for perfectly collimated mass loss. We
approximate the outflow speed as a function of heliocentric distance by mean thermal speed
vth = p8kBT / (πµmH), where µ = 18 is the molecular mass for the water-ice sublimation
scenario, mH = 1.67 × 10−27 kg is the mass of the hydrogen atom and kB = 1.38 × 10−23 J
K−1 is the Boltzmann constant. We solve for the surface temperature, T , using the energy
balance equation
(1 − A) S⊙
r2
cos ζ = ǫσT 4 + L (T ) Z (T )
(4)
in combination with the Clausius-Clapeyron relation for water ice. Here, A is the Bond
albedo, S⊙ = 1361 W m−2 is the solar constant, cos ζ is the effective projection factor for
the surface, r is expressed in AU, ǫ is the emissivity, σ = 5.67 × 10−8 W m−2 K−4 is the
Stefan-Boltzmann constant, L (T ) in J kg−1 is the latent heat of vaporization, and Z (T ) in
molecules per unit time per unit area is the gas production rate per unit area of surface.
In this study, we assume ǫ = 1, and cos ζ = 1/4, the latter corresponding to an isothermal
nucleus, while L(T ) is documented in Huebner et al.
(2006). The Bond albedos of the
active asteroids are computed according to their geometric albedos by following the method
by Bowell et al. (1989). The choice of cos ζ = 1/4 is made to remain consistent with the
isothermal assumption by Marsden et al. (1973) (but see Appendix A).
The collimation efficiency remains observationally unconstrained, although observations
showing that cometary emissions are largely sunward suggest that small values of κ are
-- 8 --
unrealistic. We choose κ ≡ 0.8 for the sake of definiteness. Combined with Equation (4), the
time-average mass-loss rate around the orbit can be numerically estimated by transforming
Equation (3) to
πρD3pA2
1 + A2
6κP
2 + A2
3
M ≃ −
Z P
0
g (r (t))
vth (r (t))
dt,
(5)
where ρ is the bulk density, D is the diameter of the body, and P is the orbital period. We
assume nominal density ρ = 103 kg m−3 for all the active asteroids, while D is extracted
from either the JPL Small-Body Database Browser or Table 2 in Jewitt et al. (2015). The
M solely from the covariance
results are listed in Table 3. We calculated the uncertainty of
matrix of Aj (j = 1, 2, 3) based upon error propagation. For cases where objects have SNR
≤ 3 for
M , we list 5σ upper limits to the values.
The upper limits to mass-loss rates inferred dynamically are consistent with, but less
stringent than, published mass-loss rates inferred from physical observations. Although A2 is
formally significant for 313P/Gibbs, large uncertainties in A1 and A3 degrade the total SNR
M is given in the table. The dynamical
to < 3, and therefore only a 5σ upper limit for its
estimate for the mass-loss rate of 324P/La Sagra (36 ± 3 kg s−1), however, exceeds values
obtained from physical observations (∼0.2 -- 4 kg s−1; Moreno et al. (2011), Hsieh et al. (2012),
Jewitt et al. (2016)) by at least an order of magnitude. Notably, while 324P/La Sagra was
active, it exhibited the highest ratio of the ejected dust mass to the nucleus mass amongst
the active asteroids currently known (Hsieh 2014), suggesting an inherently higher water-ice
content. Intriguingly, it is one of the active asteroids identified by Hsieh & Haghighipour
(2016) as a potential captured Jupiter-family comet. This is likely correlated to our finding
that 324P/La Sagra has the most significant detection in the non-gravitational acceleration.
For (3200) Phaethon, since the detection of its radial non-gravitational acceleration is likely
spurious, we only present a 5σ upper limit (< 200 kg s−1) in Table 3. This weak limit is
consistent with the perihelion value (∼3 kg s−1; Jewitt et al. 2013), as well as the average
rate needed to sustain the Geminid stream over its lifetime (Jewitt et al. 2015). In neither
case, however, is a firm physical interpretation possible, because it is not known how well
the adopted momentum-transfer law represents mass loss that may be highly stochastic in
nature.
-- 9 --
4.3. Change in Orbital Elements
The presence of a non-zero non-gravitational force results in a change of the orbit. Here
we proceed to study changes in the semimajor axis, a, and eccentricity, e, due to the non-
gravitational effect, which can be calculated by means of Gauss' form of Lagrange's planetary
equations
e sin θ
√1 − e2
P
π (cid:20)AR
P√1 − e2
2πa
a =
e =
a√1 − e2
r
(cid:21) ,
+ AT
[AR sin θ + AT (cos θ + cos E)] ,
(6)
(7)
where θ is the true anomaly, and E is the eccentric anomaly (Danby 1992). We consider
their time-average values by
g (r)
A2a√1 − e2
Z P
A2√1 − e2
g (r)(cid:20)cos θ +
Z P
2πa
dt,
π
r
0
0
¯a ≃
¯e ≃
1
e(cid:16)1 −
r
a(cid:17)(cid:21) dt,
(8)
(9)
Here we have assumed that all of the orbital elements are changing very slowly, such that
only θ-dependent functions cannot be taken out of the integral. All the terms containing
sin θ in the right-hand side of Equations (6) and (7) are eliminated thanks to the orbital
symmetry.
By substituting time t with the eccentric anomaly θ (see Appendix B), we obtain
P A2
π2a Z π
2π2a3 Z π
P A2
0
0
¯a ≃
¯e ≃
rg (r) dθ,
r2g (r)(cid:20)cos θ +
1
e(cid:16)1 −
r
a(cid:17)(cid:21) dθ,
(10)
(11)
Note that Equations (10) and (11) are only applicable to objects not in strong mean-motion
resonances with Jupiter, the most massive planet in the solar system, because the gravita-
tional influence from Jupiter is simply ignored. Indeed, none of the active asteroids are in
strong mean-motion resonances with Jupiter. We list the results in Table 3. 324P/La Sagra
-- 10 --
has the most interesting result, with astoundingly large ¯a and ¯e. The trend indicates that
its heliocentric orbit is rapidly becoming smaller and more circular. The timescale to drift
∼1 AU, if the non-gravitational effect is persistent, would be ∼105 yr. Sustained dynamical
evolution on this timescale means that we cannot be sure of the origin of this body, either
as a short-period comet trapped from the Kuiper belt or as an icy asteroid from another
part of the main-belt. On the other hand, however, its huge A2 suggests a very short active
lifetime, limited by the availability of volatiles. Using only physical observations, Jewitt et
al. (2016) reported a lifetime to mass loss of ∼105 yr and concluded that, to survive for the
expected ∼0.4 Gyr collisional lifetime, the body must lie dormant for all but 0.02 -- 0.08% of
the time. In this regard, the inferences from the orbit and from physical observations are
concordant.
4.4. Other Physical Mechanisms
We are aware that several mechanisms other than sublimation account for mass-loss
from some of the active asteroids (Jewitt et al. 2015). While the Yarkovsky effect and
the solar radiation pressure force can impart non-gravitational accelerations on an active
asteroid in a continuous manner similar to sublimation activity, non-gravitational forces
due to rotational instability and impacts obviously cannot be described by the momentum-
transfer law in the formalism by Marsden et al. (1973). In particular, mass shedding from
rotational instability is believed to be extremely stochastic, as evidenced by distinguishing
differences in morphologies between active asteroids possibly experiencing rotational instabil-
ity (311P/PANSTARRS, 331P/Gibbs, P/2010 A2, and P/2013 R3; Jewitt et al. 2015). We
should not expect any detection in non-gravitational effects for these objects, because, first,
there is no preference on directions of mass shedding, and second, astrometry from relatively
low-resolution observations normally contains larger errors in centroiding optocenters, once
there are other fragments apparently close to the primary. Indeed, we have no detections in
non-gravitational effects for the active asteroids undergoing suspected rotational instability
(see Table 1).
The momentum-transfer law by Marsden et al.
(1973) also fails for active asteroids
suffering from collision-induced mass loss, including (493) Griseldis (Tholen et al. 2015)
and (596) Scheila (Ishiguro et al. 2011a,b). The momentum-transfer law for impacts should
instead be a Dirac delta function at the time of collision. We investigate changes in the
orbital elements for these two active asteroids, considering gravity alone, by comparing the
results before and after the impact for each object. No statistically significant detection of
orbital change is made. We think that this is in agreement with Ishiguro et al. (2011a) that
-- 11 --
the impactor (∼10 m) was much smaller than (596) Scheila (∼102 km). For (493) Griseldis,
there is unfortunately no size estimate for the impactor.
4.4.1. Solar Radiation
The non-gravitational acceleration of a spherical body subjected to solar radiation pres-
sure is given by
(AR)rad =
3 (1 + A) S⊙
2cρDr2
,
(12)
where g (r) remains unchanged from Equation (2).
B) for each active asteroid. If its source is regarded as from water-ice sublimation, the cor-
where c = 3×108 m s−1 is the speed of light, and r is expressed in AU. We examine the time-
average radiation acceleration at mean heliocentric distance hri = a 4√1 − e2 (see Appendix
responding radial non-gravitational parameter is then given by (cid:16) A1(cid:17)rad ≃ (cid:0)AR(cid:1)rad /g (hri),
We present the results in Table 3, where we can see that the observed A1 is at least
an order of magnitude larger than (cid:16) A1(cid:17)rad
. It therefore suggests that either this effect is
too small among the active asteroids, or the uncertainty from the observations is too large
to enable such a detection. So far only some near-earth asteroids of ∼10 m size have been
observed to show measurable acceleration due to solar radiation pressure (e.g. Micheli et
al. 2014). Therefore, we think that the influence of the solar radiation pressure on the (much
larger) active asteroids is negligible.
4.4.2. Yarkovsky Effect
The other important physical mechanism which can give rise to a non-gravitational
Its transverse
acceleration of a sub- or kilometer-sized asteroid is the Yarkovsky effect.
acceleration is given by
(AT)Y = CY
≤ CY
ǫσT 3
cρD ∆T cos ψ
ǫσT 3
cρD ∆T
(13)
-- 12 --
where CY is a dimensionless parameter which is related to the object's shape, ∆T is the
temperature difference between the morning and evening hemispheres, and ψ is the obliquity
of the object. Thanks to the normalisation to r = 1 AU, the relationship (A2)Y ∝ D−1, where
(A2)Y is the transverse non-gravitational parameter due to the Yarkovsky effect, is then
roughly satisfied. We therefore use (A2)Y,Bennu, the transverse non-gravitational parameter
due to the Yarkovsky effect of asteroid (101955) Bennu, hitherto the most reliable and
strongest detection, as a reference to assess expected values for the active asteroids
(cid:12)(cid:12)(cid:12)(A2)Y,exp(cid:12)(cid:12)(cid:12) =(cid:12)(cid:12)(cid:12)(A2)Y,Bennu(cid:12)(cid:12)(cid:12)
DBennu
D
,
(14)
where (A2)Y,Bennu = −4.5 × 10−14 AU day−2, and DBennu = 0.49 km is Bennu's diameter
(Farnocchia et al. 2013).
The semimajor-axis drift due to the Yarkovsky effect can be computed by Equation
(10), with g(r) = r−m, where the exact value of m depends upon thermal properties of the
asteroid which are, unfortunately, poorly known. However, the choice of m has little effect
in a typical range of 2 < m < 3 in the computation (Farnocchia et al. 2013), and thus we
adopt m = 2. Consequently, the expected drift in the semimajor axis can be simplified as
P(cid:12)(cid:12)(cid:12)(A2)Y,Bennu(cid:12)(cid:12)(cid:12) DBennu
πa2 (1 − e2) D
.
(15)
(cid:12)(cid:12)(cid:12)(¯a)Y,exp(cid:12)(cid:12)(cid:12) ≃
If the non-gravitational effect of the active asteroid is purely due to the Yarkovsky effect,
the criterion ¯a .(cid:12)(cid:12)(cid:12)(¯a)Y,exp(cid:12)(cid:12)(cid:12) must be satisfied, where ¯a is listed in Table 3. By comparison,
we notice that (2201) Oljato, and (3200) Phaethon are the only two4 potential candidates
whose motions might be influenced by the Yarkovsky effect, and we proceed to calculate
their (A2)Y, by utilising the same procedures as described in Section 2. The results are
summarized in Table 4. Unfortunately, neither of the active asteroids show statistically
significant detections. We therefore conclude that no Yarkovsky effect is detected amongst
the active asteroids.
It is noteworthy that we failed to reproduce (A2)Y of (3200) Phaethon reported by
Chernetenko (2010) and Galushina et al. (2015) even though observations after 2015 were
discarded as a means to use a similar shorter observing arc. A possible explanation is that
they might have assigned too aggressive weights to some of the observations and thus the
4Active asteroid (62412) 2000 SY178 seemingly satisfies the criterion as well, but it is disqualified by the
huge uncertainty in A2 (see Table 1).
-- 13 --
uncertainty decreases while the nominal (A2)Y may increase. Instead, our finding of (A2)Y
of (3200) Phaethon is in good match with D. Farnocchia (2016, private communication).
5. SUMMARY
We examined 18 active asteroids in search of evidence for non-gravitational accelerations
caused by anisotropic mass-loss, with the following results:
1. Three active asteroids (313P/Gibbs, 324P/La Sagra and (3200) Phaethon), exhibit
non-gravitational accelerations with at least one component having formal signal-
to-noise ratio SNR > 3. We are confident in the non-gravitational detections of
313P/Gibbs and, especially, 324P/La Sagra, both kilometer-scale objects with orbital
semi-major axes near 3 AU. However, the derived non-gravitational acceleration of
(3200) Phaethon, although formally significant, is influenced by systematic uncertain-
ties of measurement, and we do not regard it as real.
2. Upper limits to the mass-loss rates implied by our non-detections of non-gravitational
acceleration are less sensitive than, but broadly consistent with, rates inferred inde-
pendently from physical observations. However, the rate inferred for 324P/La Sagra
(∼36 kg s−1) is an order of magnitude larger than values based on physical observations
(0.2 -- 4 kg s−1). The reason for this disagreement is not known, but may relate to the
poor approximation to impulsive mass loss given by the use of the non-gravitational
force law by Marsden et al. (1973).
3. The momentum-transfer law devised by Marsden et al. (1973) assumes sublimation
from an isothermal surface and is logically inconsistent with the existence of non-
gravitational acceleration (Appendix A). Anisothermal surface temperature distribu-
tions are physically more plausible and should replace the law by Marsden et al. (1973).
Except in special cases, the law proposed here (Table 5) will give similar results for
the derived non-gravitational parameters.
4. We find no evidence for radiation pressure acceleration or the Yarkovsky effect in our
sample.
We thank the anonymous referee for helpful comments and suggestions. This work used
the Find Orb code by Bill Gray, for whose assistance we are extremely grateful. We are
indebted to Aldo Vitagliano, Davide Farnocchia, and Quan-Zhi Ye for insightful discussions.
-- 14 --
We also thank all observers who submitted astrometric data to the Minor Planet Center,
except the ones who submitted really bad astrometry and thus tortured us. This work is
funded by a grant from NASA to DJ.
-- 15 --
A. THE MARSDEN MOMENTUM TRANSFER LAW
The momentum-transfer law by Marsden et al. (1973) has been widely used to calculate
non-gravitational accelerations of comets. It assumes that sublimation proceeds at a rate
appropriate for a uniform, isothermal, spherical nucleus in instantaneous equilibrium with
sunlight. However, an isothermal, spherical nucleus would sublimate isotropically, producing
no recoil force. Therefore, the law by Marsden et al. (1973) is logically inconsistent with
the presence of non-gravitational acceleration. We briefly examine the significance of this
inconsistency.
As limiting cases, we compare in Figure (1) the model by Marsden et al. (1973) (solid
black line) with three different solutions to Equation (4). Our approximation to isothermal
sublimation (labeled cos ζ = 1/4 and shown by a red dash-dot line in the figure) essentially
reproduces that by Marsden et al. (1973). Models in which sunlight heats only the day-
side of the nucleus (cos ζ = 1/2, dashed green line) and in which heat is deposited only
at the sub-solar point (cos ζ = 1, dotted blue line) both show substantially higher specific
sublimation rates at r & 2.5 AU as a result of the higher average temperatures. The revised
non-gravitational parameters for these models are listed in Table 5.
To test the effect of the differences shown in Figure (1), we computed new orbits of
selected short-period and Halley-type comets with nonzero non-gravitational effects5 using
astrometric data from the MPC Database Search with the parameters in Table 5. We found
that, even when using the two most extreme scenarios (namely, the isothermal (cos ζ =
1/4) and subsolar (cos ζ = 1) models), the derived orbital solutions and time-average non-
gravitational accelerations are unchanged, within the uncertainties. Specifically, the RMS
of best fits computed using the different momentum transfer laws of Table 5 are basically
the same. Physically, this is because the differences between the sublimation curves in
Figure (1) are significant only at r & 2.5 AU, where the momentum flux driven by water-
ice sublimation is already very low. Nevertheless, our suggestion is for future work to use
the best-fit parameters given in Table 5 for cos ζ = 1/2. This case is physically the most
plausible, since cometary nuclei are observed to sublimate primarily from the dayside (Keller
et al. 2004), and it is also logically consistent with a net force acting on the nucleus.
Of course in reality, non-gravitational effects due to mass-loss activity are strongly
dependent on, for instance, the shape, topography, spin, and thermal properties of individual
nuclei, as well as the distribution of volatiles. It is impractical to devise a model which can
5This was checked through the JPL Small-Body Database Search Engine. Only comets with >10σ
detections on non-gravitational effects were selected.
-- 16 --
universally satisfy all the cases of such complexity. Besides, little is known about the nuclei
of the majority of comets. Therefore, adopting the aforementioned simplistic model is still
appropriate and necessary for most cases.
B. DERIVATION OF TIME-AVERAGE VALUES
Let us consider a continuous function of time t which is symmetric about axes of a body's
elliptical orbit, denoted as f (t). The elliptical orbit has semimajor axis a and eccentricity
e. Now the task is to find its time-average value
¯f =
1
P Z P
0
f (t) dt,
(B1)
where P is the orbital period. Because f (t) is symmetric about the axes of the ellipse, i.e.,
f (P − t) = f (t), Equation (B1) is therefore equivalent to
¯f =
2
P Z P
2
0
f (t) dt.
(B2)
It is often the case where f is explicitly a function of true anomaly θ, i.e., f = f (θ),
and henceforth we need to find a way which connects θ and t. From orbital mechanics we
know the following relationships:
P
2π
M,
t − t0 =
M = E − sin E,
E = arccos(cid:18) e + cos θ
1 + e cos θ(cid:19) ,
(B3)
(B4)
(B5)
where M is the mean anomaly, and E is the eccentric anomaly. Differentiating both sides
from Equation (B3) to (B5) yields
dt =
P
2π
dM,
dM = (1 − cos E) dE,
dθ.
dE =
√1 − e2
1 + e cos θ
(B6)
(B7)
(B8)
-- 17 --
We then apply the chain rule to Equation (B2) and obtain
dM
dE
dt
dM
f
¯f =
=
2
P Z π
dE
dθ
dθ
(1 − e2)3/2
0
π
Z π
0
dθ
f (θ)
(1 + e cos θ)2 .
(B9)
Under polar coordinates with one of the foci at the origin, which represents the Sun,
and the other focus on the negative x-axis, the elliptical orbit is expressed by
r =
a (1 − e2)
1 + e cos θ
.
Combining Equations (B9) with (B10), we derive
¯f =
1
πa2√1 − e2 Z π
0
dθf (θ) r2.
(B10)
(B11)
In this study we need mean temperatures of the active asteroids, whose orbits are
approximately elliptic, by ignoring perturbations from other bodies and non-gravitational
effects. In accordance with Equation (4), we have f = r−2 in this scenario. Immediately, we
obtain
r2(cid:19) =
(cid:18) 1
1
a2√1 − e2
.
(B12)
The equivalent mean heliocentric distance under this definition is thereby hri = a 4√1 − e2.
Interestingly, the time-average heliocentric distance is ¯r = a (1 + e2/2), given by Equation
(B9) with f = r.
-- 18 --
REFERENCES
Baer, J., Chesley, S. R., & Matson, R. D. 2011, AJ, 141, 143
Bowell, E., Hapke, B., Domingue, D., et al. 1989, Asteroids II, eds. R. Binzel, T. Gehrels &
M. Matthews, University of Arizona Press, Tucson, p. 524
Chernetenko, Y. A. 2010, Protecting the Earth against Collisions with Asteroids and Comet
Nuclei, Proceedings of the International Conference "Asteroid-Comet Hazard 2009",
eds. A. M. Finkelstein, W. F. Huebner, & V. A. Shor (St. Petersburg: Nauka), p.289
Chesley, S. R., Ostro, S. J., Vokrouhlick´y, D., et al. 2003, Science, 302, 1739
Chesley, S. R., Kaluna, H., Kleyna, J., et al. 2010, Bulletin of the American Astronomical
Society, 42, 950
Chesley, S. R., Baer, J., & Monet, D. G. 2010, Icarus, 210, 158
Chesley, S. R., Nolan, M. C., Farnocchia, D., et al. 2012, Asteroids, Comets, Meteors 2012,
1667, 6470
Danby, J. M. A. 1992, Fundamentals of Celestial Mechanics (Richmond, VA, U.S.A.:
Willman-Bell, Inc., 2003. 2nd ed., revised & enlarged )
Farnocchia, D., Chesley, S. R., Vokrouhlick´y, D., et al. 2013, Icarus, 224, 1
Farnocchia, D., Chesley, S. R., Chodas, P. W., et al. 2014, ApJ, 790, 114
Galushina, T. Y., Ryabova, G. O., & Skripnichenko, P. V. 2015, Planet. Space Sci., 118, 296
Hsieh, H. H. 2014, Icarus, 243, 16
Hsieh, H. H., & Jewitt, D. 2006, Science, 312, 561
Hsieh, H. H., Yang, B., Haghighipour, N., et al. 2012, AJ, 143, 104
Hsieh, H. H., & Haghighipour, N. 2016, Icarus, 277, 19
Huebner, W. F., Benkhoff, J., Capria, M.-T., et al. 2006, Heat and Gas Diffusion in Comet
Nuclei, by Walter F. Huebner, Johannes Benkhoff, Maria-Theresa Capria, Angioletta
Coradini, Christina De Sanctis, Roberto Orosei, and Dina Prialnik. SR-004, June,
2006. ISBN 1608-280X. Published for The International Space Science Institute, Bern,
Switzerland, by ESA Publications Division, Noordwijk, The Netherlands, 2006.,
Hui, M.-T., & Jewitt, D. 2015, AJ, 149, 134
-- 19 --
Hui, M.-T., & Li, J. 2016, arXiv:1611.07061
Ishiguro, M., Hanayama, H., Hasegawa, S., et al. 2011, ApJ, 740, L11
Ishiguro, M., Hanayama, H., Hasegawa, S., et al. 2011, ApJ, 741, L24
Jewitt, D. 2012, AJ, 143, 66
Jewitt, D., & Li, J. 2010, AJ, 140, 1519
Jewitt, D., Li, J., & Agarwal, J. 2013, ApJ, 771, L36
Jewitt, D., Hsieh, H., & Agarwal, J. 2015, Asteroids IV, eds. P. Michel, F. DeMeo, & W.
Bottke, University of Arizona Space Science Series, 895 pp, p. 221
Jewitt, D., Agarwal, J., Weaver, H., et al. 2016, arXiv:1606.08522
Keller, H. U., Britt, D., Buratti, B. J., & Thomas, N. 2004, Comets II, eds. M. C. Festou,
H. U. Keller, & H. A. Weaver, University of Arizona Press, Tucson, 745 pp., p. 211
Levison, H. F., Terrell, D., Wiegert, P. A., Dones, L., & Duncan, M. J. 2006, Icarus, 182,
161
Li, J., & Jewitt, D. 2013, AJ, 145, 154
Marsden, B. G., Sekanina, Z., & Yeomans, D. K. 1973, AJ, 78, 211
Micheli, M., Tholen, D. J., & Elliott, G. T. 2014, ApJ, 788, L1
Moreno, F., Lara, L. M., Licandro, J., et al. 2011, ApJ, 738, L16
Nugent, C. R., Margot, J. L., Chesley, S. R., & Vokrouhlick´y, D. 2012, AJ, 144, 60
Sekanina, Z., & Kracht, R. 2015, ApJ, 801, 135
Stevens, B. L., Sarneczky, K., Wainscoat, R. J., et al. 2016, Minor Planet Electronic Circu-
lars, 2016-G72,
Tholen, D. J., Sheppard, S. S., & Trujillo, C. A. 2015, AAS/Division for Planetary Sciences
Meeting Abstracts, 47, #414.03
Tubbiolo, A. F., Bressi, T. H., Wainscoat, R. J., et al. 2015, Minor Planet Electronic Circu-
lars, 2015-X180,
Vokrouhlick´y, D., Chesley, S. R., & Matson, R. D. 2008, AJ, 135, 2336
-- 20 --
Yeomans, D. K., Chodas, P. W., Sitarski, G., Szutowicz, S., & Kr´olikowska, M. 2004, Comets
II, eds. M. C. Festou, H. U. Keller, & H. A. Weaver, University of Arizona Press,
Tucson, 745 pp., p.137
This preprint was prepared with the AAS LATEX macros v5.2.
Table 1. Non-Gravitational Parameters of Active Asteroids
Object
A1
SNR(A1)
A2
SNR(A2)
A3
SNR(A3)
Data arc
# obs†
# opp‡ RMS
(AU day−2)
−1.15 × 10−11
+5.09 × 10−10
−4.83 × 10−10
−4.18 × 10−8
−2.88 × 10−8
−1.26 × 10−10
+2.28 × 10−9
+3.27 × 10−8
−2.96 × 10−7
−1.09 × 10−7
+6.71 × 10−11
+7.53 × 10−12
+4.67 × 10−13
+6.97 × 10−12
+5.20 × 10−10
−1.76 × 10−7
−6.52 × 10−6
+1.65 × 10−6
107P
133P
176P
238P
259P
288P
311P
313P
324P
331P
493
596
2201
3200
62412
P/2010 A2
P/2012 T1
P/2013 R3
(AU day−2)
−3.56 × 10−14
+3.63 × 10−12
−1.04 × 10−11
−3.40 × 10−8
+5.17 × 10−9
+4.69 × 10−12
+3.12 × 10−11
+2.13 × 10−8
−1.47 × 10−7
+5.16 × 10−10
−2.47 × 10−12
−1.16 × 10−12
+2.95 × 10−14
−1.44 × 10−15
−1.53 × 10−14
+7.97 × 10−8
−1.06 × 10−6
+6.80 × 10−7
(AU day−2)
+1.64 × 10−11
−1.14 × 10−10
−9.12 × 10−11
+6.12 × 10−12
+1.10 × 10−8
−5.31 × 10−10
−6.36 × 10−10
−4.82 × 10−9
−3.75 × 10−8
+6.58 × 10−9
+1.74 × 10−12
−1.85 × 10−10
−3.36 × 10−12
+8.88 × 10−13
+1.02 × 10−9
−1.10 × 10−7
+2.22 × 10−7
−5.23 × 10−8
2.58
0.33
0.42
2.13
0.70
0.09
2.23
4.45
10.50
0.87
1.80
1.75
2.29
0.92
< 1%
2.21
1.58
1.40
2.03
2.62
2.64
1.13
2.97
0.20
1.85
1.75
10.46
2.24
0.73
0.22
0.15
3.40
0.83
1.56
1.42
1.04
1.97
0.33
0.18
< 1%
2.61
1.38
1.10
1.83
7.41
0.96
0.01
2.14
0.32
0.59
1.08
1.34
1.27
2.19
1949 -- 2016
1979 -- 2016
1999 -- 2016
2005 -- 2011
2008 -- 2012
2000 -- 2015
2005 -- 2015
2003 -- 2014
2010 -- 2015
2004 -- 2015
1902 -- 2016
1908 -- 2016
1931 -- 2015
1983 -- 2016
1999 -- 2016
2010 -- 2012
2012 -- 2013
2013 -- 2014
909 (17)
716 (13)
568 (2)
141 (0)
40 (6)
160 (0)
158 (3)
94 (3)
421 (2)
148 (10)
1388 (29)
3418 (71)
823 (23)
3161 (60)
737 (2)
127 (95)
165 (1)
316 (5)
18
18
14
4
4
9
5
3
4
6
27
41
25
30
13
2
1
1
(′′)
0.57
0.50
0.48
0.59
0.73
0.52
0.45
0.63
0.48
0.86
0.51
0.40
0.51
0.46
0.54
1.23
0.45
0.63
--
2
1
--
†Total number of observations of all types (optical and radar) used in fit. Number of discarded data bracketed.
‡Number of observed oppositions
Note. -- The non-gravitational parameters are calculated based on the isothermal water-ice sublimation model devised by Marsden et al. (1973).
SNR(Aj ) (j = 1, 2, 3) is the ratio of Aj over its 1σ uncertainty. All of the astrometric observations were retrieved on 2016 July 14 -- 15.
Table 2. Non-Gravitational Parameters of Some Moderate-Size Asteroids
Object
A1
SNR(A1)
A2
SNR(A2)
A3
SNR(A3)
Data arc
# obs†
# opp‡ RMS
(AU day−2)
−2.73 × 10−11
−3.29 × 10−12
+1.86 × 10−11
+6.59 × 10−12
−7.85 × 10−12
−9.26 × 10−11
+2.52 × 10−11
−6.45 × 10−12
+3.16 × 10−11
+6.32 × 10−13
−1.36 × 10−11
+4.64 × 10−11
3818
7916
9344
11313
13426
16392
18333
20099
20293
23059
25343
26662
(AU day−2)
+2.31 × 10−13
+1.72 × 10−13
+1.93 × 10−12
−1.48 × 10−12
−2.37 × 10−13
−2.37 × 10−13
−1.20 × 10−13
−5.11 × 10−12
−9.05 × 10−13
−3.62 × 10−13
−1.66 × 10−12
−4.82 × 10−13
1.65
0.23
0.58
0.09
0.45
1.84
1.02
0.05
2.09
0.05
0.50
1.08
(AU day−2)
+3.19 × 10−11
+2.56 × 10−11
+1.12 × 10−10
+1.67 × 10−10
+1.97 × 10−11
+5.85 × 10−11
+3.83 × 10−11
+2.97 × 10−11
+6.33 × 10−11
+1.97 × 10−11
+7.86 × 10−11
+7.31 × 10−11
0.62
0.61
2.59
1.69
0.86
0.13
0.09
0.51
1.66
0.76
2.54
0.62
1.16
1.03
2.45
1.90
0.71
0.57
0.71
0.14
1.94
0.80
1.95
2.09
1979 -- 2015
1978 -- 2015
1991 -- 2016
1976 -- 2016
1975 -- 2015
1977 -- 2016
1987 -- 2016
1991 -- 2015
1980 -- 2016
1991 -- 2016
1992 -- 2015
1974 -- 2015
1166 (16)
1080 (5)
1222 (6)
1219 (3)
792 (2)
1085 (2)
1100 (4)
852 (1)
1099 (5)
1240 (1)
866 (6)
636 (1)
20
18
16
18
14
19
16
17
15
15
16
17
(′′)
0.49
0.53
0.54
0.52
0.54
0.50
0.54
0.49
0.52
0.47
0.56
0.56
†Total number of observations of all types (optical and radar) used in fit. Number of discarded data bracketed.
‡Number of observed oppositions
--
2
2
--
Note. -- All of the asteroids have diameters ∼10 km. The non-gravitational parameters are calculated based on the isothermal water-ice sublimation
model devised by Marsden et al. (1973). All of the astrometric observations were retrieved on 2016 July 14 -- 15.
-- 23 --
Table 3. Physical and Derived Properties
Object
D(1)
A(2)
(km)
− M (3)
(kg s−1)
(cid:0)AR(cid:1)rad
(AU day−2)
(4)
(cid:16)eA1(cid:17)rad
(AU day−2)
(5)
¯a(6)
(AU yr−1)
¯e(7)
(yr−1)
< 0.5
3.5
3.8
4.0
0.8
0.6
3
107P
133P
176P
238P
259P
288P
311P
313P
324P
331P
493
596
2201
3200
62412
P/2010 A2
P/2012 T1
P/2013 R3 < 0.4
1.0
1.1
1.8
46.4
113.3
1.8
5.1
7.8
0.12
2.4
0.02
0.02
0.02
0.02
0.02
0.02
0.11
0.02
0.02
0.02
0.02
0.01
0.17
0.04
0.03
0.04
0.02
0.02
< 5
< 4
< 5
< 13
< 32
< 8
< 1
< 12
36 ± 3
. 77
. 103
. 105
< 2
< 200
< 70
< 1
. 104
< 141
1.82 × 10−14
9.26 × 10−15
8.68 × 10−15
4.48 × 10−14
8.25 × 10−14
1.26 × 10−14
3.59 × 10−14
3.31 × 10−14
2.13 × 10−14
7.81 × 10−16
3.60 × 10−16
6.68 × 10−14
9.36 × 10−14
4.51 × 10−15
5.65 × 10−13
1.49 × 10−14
5.06 × 10−13 −1.9 × 10−9 −2.7 × 10−10
1.14 × 10−11 +3.2 × 10−9 +6.5 × 10−10
1.24 × 10−11 −1.2 × 10−8
−2.5 × 10−9
4.76 × 10−11 −9.9 × 10−5
−2.0 × 10−5
8.35 × 10−12 +6.7 × 10−5
+1.3 × 10−5
+2.0 × 10−9
8.14 × 10−12 +9.5 × 10−9
> 1.59 × 10−13 > 2.56 × 10−12 +3.1 × 10−7
+4.1 × 10−8
3.76 × 10−11 +5.4 × 10−5
+1.1 × 10−5
2.95 × 10−11 −1.4 × 10−4
−2.8 × 10−5
1.28 × 10−11 +2.1 × 10−7
+2.0 × 10−8
7.84 × 10−13 −2.7 × 10−9 −5.5 × 10−10
1.30 × 10−13 −2.2 × 10−9 −4.5 × 10−10
3.81 × 10−13 +2.8 × 10−9 +3.8 × 10−10
6.66 × 10−14 −9.4 × 10−10 −2.6 × 10−10
5.87 × 10−12 −4.6 × 10−12 −7.3 × 10−13
1.29 × 10−11 +6.5 × 10−4
+9.8 × 10−5
−5.2 × 10−4
1.57 × 10−11 −2.5 × 10−3
> 9.77 × 10−14 > 5.09 × 10−11 +3.3 × 10−3
+6.7 × 10−4
(1)Diameter
(2)Bond albedo
(3)Time-average mass-loss rate estimated from Equation (5)
(4)Computed non-gravitational acceleration due to the solar radiation force
(5)Radial non-gravitational parameter due to the solar radiation force but computed with the momentum-
transfer law by Marsden et al. (1973)
(6)Time-average drift in semimajor axis
(7)Time-average drift in eccentricity
Note. -- The significance levels of an orbital drift in a and e are predominantly determined by the ones
of the non-gravitational parameters, which are the most uncertain parameters compared to the rest orbital
elements. See Equations (10) and (11). Therefore, the SNRs of ¯a and ¯e are both given by SNR(A2), listed
in Table 1.
-- 24 --
Table 4. Transverse Non-Gravitational Parameters Due to the Yarkovsky Effect
Object
(cid:12)(cid:12)(cid:12)(A2)Y,exp(cid:12)(cid:12)(cid:12)†
(AU day−2)
(A2)Y
‡
(AU day−2)
Data arc
# obs∗ # opp⋆ RMS
(′′)
0.51
0.46
2201
3200
1.2 × 10−14
4.4 × 10−15
(+2.89 ± 1.28) × 10−14
(−1.39 ± 1.56) × 10−15
1931 -- 2015
1983 -- 2016
824 (22)
3161 (60)
25
30
†Value of expected transverse non-gravitational parameter due to the Yarkovsky effect
estimated from the one of (101955) Bennu through Equation (14).
‡Transverse non-gravitational parameter due to the Yarkovsky effect computed from orbit
determination.
∗Total number of observations of all types (optical and radar) used in fit. Number of
discarded data bracketed.
⋆Number of observed oppositions.
Note. -- The same technique as used for obtaining the non-gravitational parameters in
Table 1 is applied, with the modified momentum-transfer law g(r) = r−2.
Table 5. Parameters in the Momentum-Transfer Law
Parameter
cos ζ = 1/4
(Isothermal)
cos ζ = 1/2
(Hemispherical)
cos ζ = 1
(Subsolar)
Unit
α
m
n
k
r0
0.1258295
2.13294
5.30728
4.19724
2.67110
0.0337694
2.08782
4.04051
11.4543
5.10588
0.0003321
2.04680
3.06682
2752.35
50.4755
--
--
--
--
AU
Note. -- Each least-squares fit was performed for heliocentric dis-
tance r ≤ 5 AU, beyond which the contribution from the water-ice
sublimation is negligible. See Figure 1 for comparison.
-- 25 --
Fig. 1. -- Comparison of our best fits in the formalism by Equation (2) for three different
sublimation scenarios, i.e., cos ζ = 1/4 (isothermal sublimation), 1/2, and 1 (subsolar), and
the best fit by Marsden et al. (1973). The actual normalised water-ice sublimation functions
are indistinguishable from our best fits correspondingly, were they plotted in the figure, and
therefore are omitted. Different fits are discriminated by line styles.
|
1504.04379 | 2 | 1504 | 2015-09-25T02:52:58 | The K2-ESPRINT Project. I. Discovery of the Disintegrating Rocky Planet K2-22b with a Cometary Head and Leading Tail | [
"astro-ph.EP"
] | We present the discovery of a transiting exoplanet candidate in the K2 Field-1 with an orbital period of 9.1457 hr: K2-22b. The highly variable transit depths, ranging from $\sim$0\% to 1.3\%, are suggestive of a planet that is disintegrating via the emission of dusty effluents. We characterize the host star as an M-dwarf with $T_{\rm eff} \simeq 3800$ K. We have obtained ground-based transit measurements with several 1-m class telescopes and with the GTC. These observations (1) improve the transit ephemeris; (2) confirm the variable nature of the transit depths; (3) indicate variations in the transit shapes; and (4) demonstrate clearly that at least on one occasion the transit depths were significantly wavelength dependent. The latter three effects tend to indicate extinction of starlight by dust rather than by any combination of solid bodies. The K2 observations yield a folded light curve with lower time resolution but with substantially better statistical precision compared with the ground-based observations. We detect a significant "bump" just after the transit egress, and a less significant bump just prior to transit ingress. We interpret these bumps in the context of a planet that is not only likely streaming a dust tail behind it, but also has a more prominent leading dust trail that precedes it. This effect is modeled in terms of dust grains that can escape to beyond the planet's Hill sphere and effectively undergo `Roche lobe overflow,' even though the planet's surface is likely underfilling its Roche lobe by a factor of 2. | astro-ph.EP | astro-ph |
ACCEPTED FOR PUBLICATION IN THE Astrophysical Journal, 2015 SEPTEMBER 1
Preprint typeset using LATEX style emulateapj v. 5/2/11
THE K2-ESPRINT PROJECT. I. DISCOVERY OF THE DISINTEGRATING ROCKY PLANET K2-22B
WITH A COMETARY HEAD AND LEADING TAIL.
R. SANCHIS-OJEDA1, 2, S. RAPPAPORT3, E. PALLÈ4, 5, L. DELREZ6, J. DEVORE7, D. GANDOLFI8,9,
A. FUKUI10, I. RIBAS11, K. G. STASSUN12,13, S. ALBRECHT14, F. DAI3, E. GAIDOS15, M. GILLON5,
T. HIRANO16, M. HOLMAN17, A. W. HOWARD18, H. ISAACSON1, E. JEHIN6, M. KUZUHARA16,
A. W. MANN19, 20, G. W. MARCY1, P. A. MILES-PÁEZ4, 5, P. MONTAÑÉS-RODRÍGUEZ4, 5, F. MURGAS21, 22,
N. NARITA23, 24, 25, G. NOWAK4, 5, M. ONITSUKA23, 24, M. PAEGERT12, V. VAN EYLEN14, J. N. WINN3, L. YU3.
Accepted for publication in the Astrophysical Journal, 2015 September 1
ABSTRACT
We present the discovery of a transiting exoplanet candidate in the K2 Field-1 with an orbital period of
9.1457 hr: K2-22b. The highly variable transit depths, ranging from ∼0% to 1.3%, are suggestive of a planet
that is disintegrating via the emission of dusty effluents. We characterize the host star as an M-dwarf with
Teff ≃ 3800 K. We have obtained ground-based transit measurements with several 1-m class telescopes and
with the GTC. These observations (1) improve the transit ephemeris; (2) confirm the variable nature of the
transit depths; (3) indicate variations in the transit shapes; and (4) demonstrate clearly that at least on one
occasion the transit depths were significantly wavelength dependent. The latter three effects tend to indicate
extinction of starlight by dust rather than by any combination of solid bodies. The K2 observations yield a
folded light curve with lower time resolution but with substantially better statistical precision compared with
the ground-based observations. We detect a significant “bump” just after the transit egress, and a less significant
bump just prior to transit ingress. We interpret these bumps in the context of a planet that is not only likely
streaming a dust tail behind it, but also has a more prominent leading dust trail that precedes it. This effect
is modeled in terms of dust grains that can escape to beyond the planet’s Hill sphere and effectively undergo
‘Roche lobe overflow,’ even though the planet’s surface is likely underfilling its Roche lobe by a factor of 2.
Subject headings: planetary systems—planets and satellites: detection, atmospheres
1 Department of Astronomy, University of California, Berkeley, CA
94720; [email protected]
2 NASA Sagan Fellow
3 Department of Physics, and Kavli Institute for Astrophysics and Space
Research, Massachusetts Institute of Technology, Cambridge, MA 02139,
USA, [email protected]
4 Instituto de Astrofísica de Canarias (IAC), 38205 La Laguna, Tenerife,
Spain
5 Departamento de Astrofísica, Universidad de La Laguna (ULL),
38206 La Laguna, Tenerife, Spain
6 Août 17, B-4000 Liège, Belgium
6 Institut d’Astrophysique et Géophysique, Université de Liège, allée du
7 Visidyne, Inc., 111 South Bedford St., Suite 103, Burlington, MA
01803, USA; [email protected]
8 Dipartimento di Fisica, Universitá di Torino, via P. Giuria 1, I-10125,
Torino, Italy
9 Landessternwarte Königstuhl, Zentrum für Astronomie der Universität
Heidelberg, Königstuhl 12, D-69117 Heidelberg, Germany
10 Okayama Astrophysical Observatory, National Astronomical Obser-
vatory of Japan, Asakuchi, Okayama 719-0232, Japan
11 Institut de Ciències de l’Espai (CSIC-IEEC), Campus UAB, Facultat
de Ciències, Torre C5, parell, 2a pl., E-08193 Bellaterra, Spain
12 Vanderbilt University, Nashville, TN 37235
13 Fisk University, Nashville, TN 37208
14 Stellar Astrophysics Centre, Department of Physics and Astronomy,
Aarhus University, Ny Munkegade 120, DK-8000 Aarhus C, Denmark
15 Department of Geology & Geophysics, University of Hawaii, 1680
East-West Road, Honolulu, HI 96822, USA and Visiting Astronomer at
the Infrared Telescope Facility at the University of Hawaii.
16 Department of Earth and Planetary Sciences, Tokyo Institute of Tech-
nology, 2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan
17 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cam-
bridge, MA 02138
18 Institute for Astronomy, University of Hawaii, 2680 Woodlawn
Drive, Honolulu, HI 96822, USA
19 Harlan J. Smith Fellow
20 Department of Astronomy, The University of Texas at Austin, Austin,
TX 78712, USA
21 Univ. Grenoble Alpes,IPAG, F-38000 Grenoble, France
22 CNRS, IPAG, F-38000 Grenoble, France
23 National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka,
Tokyo 181-8588, Japan
24 SOKENDAI (The Graduate University for Advanced Studies), 2-21-
1 Osawa, Mitaka, Tokyo 181-8588, Japan
25 Astrobiology Center, National Institutes of Natural Sciences, 2-21-1
Osawa, Mitaka, Tokyo 181-8588, Japan
2
Sanchis-Ojeda et al. 2015
1. INTRODUCTION
The Kepler mission (Borucki et al. 2010) has revolutionized
the field of exoplanets, with some 4000 planet candidates dis-
covered to date (Mullally et al. 2015), of which at least 1000
have been confirmed (Lissauer et al. 2014, Rowe et al. 2014).
With the original objective of discovering Earth-size planets
in the habitable zone of their host stars, the telescope was
bound to also improve our understanding of close-in rocky
planets (Jackson et al. 2009; Schlaufman et al. 2010). In-
deed, the first Kepler rocky planet, Kepler-10b, had an orbital
period of only 20 hr (Batalha et al. 2011). The smallest planet
with a well measured mass and radius, Kepler-78b, also has a
very short orbital period of 8.5 hr (Sanchis-Ojeda et al. 2013),
which was instrumental in measuring its small mass of 1.7
Earth masses (Howard et al. 2013, Pepe et al. 2013). In spite
of the falloff in the numbers of Kepler exoplanet candidates
at short periods, there are 106 well vetted candidates with or-
bital periods shorter than one day (hereafter “USPs”; Sanchis-
Ojeda et al. 2014), and most of them seem to be smaller than
twice the size of Earth.
Not included among the above lists are two special transit-
ing exoplanets that are thought to be disintegrating via dusty
effluents (Rappaport et al. 2012; Rappaport et al. 2014). In
both cases it is inferred that the planets are trailed by a dust
tail whose dynamics are influenced by radiation pressure on
the dust grains. This leads to transit profiles characterized by
a pronounced depression in flux after the planet has moved
off of the stellar disk (i.e., a post-transit depression). In the
case of KIC 12557548b (Rapapport et al. 2012; hereafter
‘KIC 1255b’) the transit depths range from ∼1.2% down to
. 0.1% in an highly erratic manner, while for KOI 2700b
(KIC 8639908; Rappaport et al. 2014) the transit depths are
observed to be slowly decreasing in depth over the course of
the fours years of Kepler observations. The fact that these
‘disintegrating’ planets are relatively rare (2 of 4000 Kepler
planets) is likely due to the conditions required for their ex-
istence and detection, namely high surface equilibrium tem-
peratures and very low surface gravity, and a possibly short
disintegration lifetime of only 10-100 Myr (see, e.g., Rappa-
port et al. 2012; Perez-Becker & Chiang 2013).
The main Kepler mission had an abrupt ending when two
reactions wheels failed by March 2013. The reaction wheels
are very important to maintain the telescope pointing in a
given direction, and the telescope could no longer point to-
ward the original Kepler field. The problem was partially by-
passed by designing a new mission, called “K2”, in which the
telescope would point toward a different field of view along
the ecliptic plane every three months (Howell et al. 2014);
the spacecraft stability is improved by equalizing the Sun’s
radiation pressure forces on the solar panels. The unfortunate
demise of the reaction wheels that put an end to the main mis-
sion, also opened the possibility for new discoveries of planets
orbiting brighter stars since thousands of new bright stars are
observed in each field.
In the short lifespan of this new mission, there have been
several papers describing techniques to produce light curves
(Vanderburg & Johnson 2014; Aigrain et al. 2015; Foreman-
Mackey et al. 2015), and planet discoveries like a super-
Earth transiting a bright host star (HIP 116454, Vanderburg
et al. 2015), a triple planet system orbiting a bright M-dwarf
(K2-3, Crossfield et al. 2015), and a pair of gas giants near a
3:2 mean motion resonance (EPIC 201505350, Armstrong et
al. 2015b), with almost 20 confirmed K2 planets discovered
to date (Montet et al. 2015). There have also been several
catalogs of variable stars and eclipsing binaries (Armstrong
et al. 2015a; LaCourse et al. 2015). This paper is the first in
a series describing our discoveries using the K2 public data
releases. The name of the project, “ESPRINT”, stands for
"Equipo de Seguimiento de Planetas Rocosos INterpretando
sus Tránsitos", which in English means "Follow-up team of
rocky planets via the interpretation of their transits".
In this work we focus on the surprising discovery of another
one of these candidate disintegrating planets, this one in the
K2 Field 1 which contains only 21,647 target stars (close to
an order of magnitude fewer than in the prime Kepler field).
Even more impressive, this particular short-period exoplanet
appears to have a dominant leading dust tail and possibly an
additional trailing one, a phenomenon not seen before in as-
trophysics. The paper is organized as follows. In Section 2
we summarize the observations taken with the K2 mission
and describe how this particular object was found. In Sec-
tion 3 we describe the variable transit depths, the timing anal-
ysis, and the unusual transit profile that cannot be explained
by a solid body. We present and discuss 15 transit measure-
ments that were made in follow-up ground-based observations
in Section 4. We analyze the properties of the host star based
on a number of ground-based imaging and spectral observa-
tions in Section 5. In Section 6 we set significant constraints
on the radial velocity variations in the host star. We discuss
the wavelength dependence of the transits observed with the
GTC in Section 7. We summarize why the host of the transits
is the bright target star and not its much fainter companion
in Section 8. In Section 9 we interpret all the observations
in terms of a model in which the planet is disintegrating, and
discuss why the different characteristics and environment of
K2-22b could lead to a dominant leading dust tail. Finally, in
Section 10 we present a summary and conclusions and point
toward new lines of research that could improve our under-
standing of how disintegrating planets form and evolve.
2. K2 DATA PROCESSING
The target star, with EPIC number 201637175 (from now
on named K2-22), was selected as one of the 21,647 stars in
Field 1 to be observed in the long cadence mode of the K2
mission (Howell et al. 2014). During the period from 2014
May 30 to 2014 August 20, a total of 3877 images of 15 × 15
pixels were recorded by the Kepler telescope, with a typi-
cal cadence of 29.42 minutes. The data were sent to NASA
Ames, subsequently calibrated, including cosmic ray removal
(Howell et al. 2014), and uploaded to the public K2 MAST
archive in late 2014 December. The data were then down-
loaded from the MAST archive and utilized for the analysis
presented in this work.
The discovery of K2-22b is part of the larger ESPRINT col-
laboration to detect and quickly characterize interesting plan-
etary systems discovered using the K2 public data. In this sec-
tion we highlight the way in which we produce light curves for
all the observed stars, how this object was identified as part of
the survey, and how we produced a better quality light curve
for this particular object once the transits had been detected.
2.1. ESPRINT Photometric Pipeline
Our photometric pipeline follows the steps of similar ef-
forts published to date (Vanderburg & Johnson 2014; Cross-
field et al. 2015; Aigrain et al. 2015; Foreman-Mackey et al.
2015; Lund et al. 2015) that describe how to efficiently ex-
tract light curves from the calibrated pixel level data archived
K2-22b
3
on MAST. The ingredients to generate the light curves, with
our own choice described, are:
• Aperture selection: Our apertures have irregular
shapes, which are based on the amount of light that
a certain pixel receives above the background level.
We selected this type of aperture to capture as much
light as possible while reducing the number of pixels
used, which in turn reduces the noise induced by a large
background correction. Based on experiments we car-
ried out on the engineering data release, selected pixels
must be 30% higher than the background level (esti-
mated from an outlier corrected median of all the pix-
els in the image) in 50% of the images in the case of
a star brighter than Kepler magnitude Kp = 11.5. For
stars fainter than magnitude Kp = 14, the selected pix-
els must be 4% higher than the background. For stars of
intermediate magnitudes, a linear interpolation of these
two thresholds is used. A simple algorithm groups con-
tiguous pixels in different apertures , and the target star
aperture is selected to be the one that contains the tar-
get star pixel position (obtained from the FITS head-
ers). This type of aperture is similar to the ones used by
Lund et al. (2015), and quite different from the circular
apertures used in many of the other pipelines.
• Thruster event removal: As highlighted in Vanderburg
& Johnson (2014), every 6 hours the telescope rolls to
maintain the targets on the defined set of pixels that are
downloaded, in what is known as a ‘thruster event’.
We recognized these events by calculating the cen-
troid motion of a particularly well behaved star (EPIC
201918073), and selected those moments where the po-
sition of the star jumps much more than usual (in the
case of these stars, 0.1 pixels in the x direction). The
images obtained during thruster events are removed
from the analysis.
• Data slicing: We split our dataset into eleven differ-
ent segments chosen to have a length of approximately
7 days, but also to contain an integer number of tele-
scope roll cycles. The first segment and the one after
a large gap (in the middle of the dataset) are not used
in the global search since they are poorly behaved in
some cases, with systematic effects induced by thermal
changes that appear after reorienting the telescope.
• Systematics removal: We calculate the centroid po-
sitions and obtain a fourth-order polynomial that de-
scribes the movement of the star in the x and y coordi-
nates. This polynomial fit is used to determine a new set
of coordinates, in which the star moves only along one
direction. The fluxes are then decorrelated first against
time and then against this moving coordinate, using a
fourth-order polynomial in each case. This process is
repeated 3 times, and the results are very robust against
problems caused by the presence of low-frequency as-
trophysical sources of noise (see Vanderburg & John-
son 2014 for a more detailed description on how this
process works).
This recipe was followed to generate the light curves of the
21,647 stars in Field-1. Among them, the light curve for EPIC
20163717 can be seen in the top panel of Fig. 1. It is interest-
ing to note that since our apertures depend on the amount of
background light, which is increasing over the course of the
observations, the number of pixels in the aperture decreases
with time. This is a desired effect, since it tends to balance
the increase of background light in the aperture by reducing
the number of pixels, and therefore changes in the flux scatter
are less severe.
These light curves are generally analyzed using two differ-
ent search algorithms: a more standard BLS routine (Kovács
et al. 2002; Jenkins et al. 2010; Ofir 2014) to search for plan-
ets with orbital periods longer than 1 day, and a more special-
ized FFT pipeline used to detect planets with orbital periods
shorter than 1-2 days (Sanchis-Ojeda et al. 2014). In this case,
due to the short orbital period of the signal, we describe only
the FFT search.
2.2. Detection via the FFT technique
Our target star K2-22 was detected as part of our search
for ultra-short period planets in the K2 Field-1 dataset using
the FFT technique. The routine automatically identifies those
objects for which a main frequency and at least one harmonic
can be distinguished above the level noise in the FFT power
spectrum (see Sanchis-Ojeda et al. 2014 for details). A total
of 2628 objects were identified in that way, but a large fraction
of them were caused by improper corrections of the 6-hour
roll of the Kepler telescope. These false detections are easy
to remove since their main frequency is always related to the
fundamental roll frequency of 4.08 rolls per day, although this
simple removal clearly affects the completeness of our search.
A total of 390 objects were selected for visual inspection, and
among them K2-22 was selected as the most promising ultra-
short orbital period planet candidate.
The FFT by which this object was discovered is shown in
Fig. 2. Note the prominent peak at 2.62 cycles/day which is
the base frequency corresponding to the 9.1457-hour period,
as well as the next 8 higher harmonics which lie below the
Nyquist limit. The overall slowly decaying Fourier ampli-
tudes with harmonic number is characteristic of short-period
planet transits.
2.3. Individualized aperture photometry
Our method for generating the light curves relies on a one-
to-one relationship between the raw flux counts and the posi-
tion of the star on the CCD chip. Any source of astrophysical
variability could distort this relationship, and this is the case
for both stellar activity induced signals and transits. A closer
inspection of the raw light curve of K2-22 shows long-term
trends that do not correlate with the centroid motion, and are
likely due to the slow rotation of the host star. These trends
are removed automatically as we fit a fourth-order polynomial
in time to each of the 7-day segments, which effectively re-
moves variability on scales longer than approximately 2 days
(see upper panel of Fig. 1).
In order to remove the effects of the transits, we first folded
the original light curve given the period obtained from the
FFT, after which we identified those orbital phases where no
transit is expected. We then ran our photometric pipeline
again, using the same aperture (see Fig. 3) but only using the
out-of-transit flux measurements to find the best fit polynomi-
als to correct for both temporal and telescope motion varia-
tions. This process reduced the photometric scatter, and en-
couraged us to try different approaches to continue improving
the light curve. We tried different combinations of polynomial
orders and also different approaches to defining the apertures,
4
Sanchis-Ojeda et al. 2015
FIG. 2.— Discovery Fourier transform of the flux data showing strong peaks
at the 9.147-hour period and all 8 harmonics that are below the Nyquist limit-
ing frequency. The arrows mark two additional harmonics aliased around the
Nyquist limit.
FIG. 1.— Lightcurve of K2-22 in time bins equal to the half-hour Kepler
long-cadence sampling time. Top panel: Light curve of the form used in our
global search for USP planets. Middle panel: Light curve processed with a
modified algorithm that better preserves stellar activity. Lower panel: Au-
tocorrelation function, with a vertical line representing the inferred rotation
period of the star.
but none of them improved the quality of the light curve (see
top panel of Fig. 1). After all the corrections, the final scatter
per 30 minute cadence is 650 ppm, which is near the mean un-
certainty obtained with our photometric pipeline for the typi-
cal K2 15th magnitude star.
We also tried to produce a light curve in which other astro-
physical signals would be preserved (e.g., starspot rotation).
During the process of detrending each of the 11 segments of
data, we saved the coefficients of the fourth-order polynomial
in time, and used them to reconstruct the signal again after
removing the centroid motion artifact. Since the aperture is
FIG. 3.— K2 image of the object K2-22. A customized aperture is defined
based on the amount of light of each pixel, and level of background light. The
blue star represents the expected position of the target star, and the Kepler
magnitude obtained from the EPIC catalog.
individually defined for each segment, we had to adjust the
mean flux level of each segment to create a continuous light
curve. This astrophysically more accurate light curve is also
shown in Fig. 1, and exhibits a clear signal of starspots with a
rotation that could either be 7-8 days or twice this value. The
shorter quasi-periodicity would typically arise when the star
has two active longitudes separated by 180 degrees in lon-
gitude. We used an autocorrelation function to confirm this
suspicion (see lower panel of Fig. 1), and measured a rota-
tion period of 15.3 days, following the techniques described
in McQuillan et al. (2013).
3. TRANSIT PARAMETERS AND DEPTH VARIABILITY
In this section we describe the transits of K2-22b as ob-
served by K2, with particular emphasis on the characteristics
that deviate from the transits of a more typical planet.
K2-22b
5
those cases, we were left with a reduced sample of 60 well
detected transits. In addition to the best fit transit time, we es-
timated the formal uncertainty by finding the interval of times
where the standard χ2 function is within 1 of its minimum
value, where a constant value of 650 ppm is used for the un-
certainties in the flux. The transit times (including uncertain-
ties) obtained in this manner are fit to a linear function, and
the O- C (‘observed minus calculated’) residuals of that fit are
displayed in Fig. 5. A clear excess of scatter (above the formal
statistical uncertainties) is detected, with a best-fit standard χ2
of 480 for 60 transit times. Formal uncertainties, however, un-
derestimate the true uncertainties when the model parameters
are correlated, as could be the case here since we have re-
moved a local linear trend for each transit which is known to
be correlated with the transit time. We further examine this
excess scatter below with simulations of the data train.
We now use the new orbital ephemeris to fix the time of
each transit, and repeat the process but now fitting for the
transit depths. These transit depths are significantly differ-
ent from one transit to the next, and the depths range from a
maximum of 1.3% to 0.27% which is close to the photometric
detectability limit of our time series for individual flux points.
We can also see in Fig. 5 that the transit depths do not strongly
correlate with the O - C timing residuals. These erratic tran-
sit depth variations are quite reminiscent of those exhibited
by KIC 1255b (Rappaport et al. 2012; Croll et al. 2014), and
cover much the same range in depths.
We checked for periodicities in the measured transit depths
and timing residuals using a Lomb-Scargle periodogram, but
no significant peak was detected with a false alarm probability
smaller than 1%. We also attempted to constrain the true size
of the planet by studying the shallowest K2 transits. We revis-
ited the discarded transits, and selected the 6 shallowest cases
in which the transit observations were complete (no thruster
events). The mean depth of these transits is 0.14 ± 0.03%,
which, given the radius of the star (see section 5.2), translates
into an upper bound of 2.5 ± 0.4 R⊕ on the planet radius.
Even though the transit depth variations described above
are later confirmed by ground-based observations, there is a
potential concern that the variability could have been caused
by the relatively comparable transit and sampling timescales.
In this same regard, it is also possible that some or all of the
excess variations in the O - C scatter (see Fig. 5) over those
expected from statistical fluctuations are due to the relatively
short transit duration compared with the LC integration time.
We have therefore carried out extensive numerical simulations
of these effects.
The simulations of the depth and transit-time variations are
based on a model that has a simple box transit profile which
is integrated over the 30 minute LC time, and includes a sinu-
soid (and its first harmonic) to represent the rotating starspot
activity. The amplitude of the sinusoids and the rotation pe-
riod are fixed at 1% and 15 days, respectively. The duration of
the model transit is fixed at 50 min (see Sect. 4). The model
is evaluated at the same times as the K2 observations, and
only the same 60 transit windows are analyzed to ensure that
the simulations represent a similar dataset to the one used in
this paper. There are a total of four different numerical ex-
periments, each one repeated many times, in which we either
include starspots or not, and we either have a constant transit
depth or depths drawn from a Gaussian distribution of a given
variance, but always with a mean of 0.5%. In all cases a fixed
orbital ephemeris is used. White noise of 700 ppm per LC
FIG. 4.— Zoomed version of Fig. 1 where the variability of the transit
depths in much easier to see. Each vertical red line represents an expected
transit time.
3.1. Individual transit times and depths
The top panel of Fig. 1 shows the full data set from the
K2 Field-1 observations covering an interval of ∼80 days. It
is apparent that there are sharp dips in intensity whose depths
are highly and erratically variable. In Fig. 4 we can see a zoom
in on two weeks of observations. The individual transits are
now quite apparent, and the depth variations are dramatically
evident.
In order to analyze these variations, we first folded the light
curve with the period obtained from the FFT, after removing
any long-period signals. We did this by fitting for local lin-
ear trends using a total of 5 hours of observations right before
and after each transit. We then fit the folded light curve with
a simple idealized transit profile comprised of 3 straight-line
segments: a flat bottom and sloping ingress and egress with
the same slope magnitudes, hereafter referred to as the “three-
segment (symmetric) model”. This is very similar to the sim-
pler ‘box model’, but with non-zero ingress and egress times.
We forced the fit to have a duration of at least 5 minutes on
the lowest part of the transit, to make sure that the final fit did
not look like a triangle. This fit gave a mean transit depth of
0.5%, and a total duration of 1 hour and 10 minutes. Given
that the cadence of the observations is approximately 30 min-
utes, the real mean duration of the K2 transits must be close
to 40 minutes, in agreement with what is expected of an USP
planet orbiting an M-dwarf. This is confirmed by our follow-
up ground-based observations with their better temporal reso-
lution (see Sect. 4).
With the mean transit profile and a good estimate for the or-
bital ephemeris, we fit each of the 190 individual K2 transits
with the same three-segment model, allowing only the tran-
sit time to vary. In the process, we evaluated the robustness
of the detection of each of the individual transits. In some
cases, the depth is so low that the transit cannot be detected,
whereas on other occasions the transit occurs during a thruster
event, so there are no data. In rare cases, the transit consists
of only a single flux point, which does not allow for a clear
determination of the transit time or depth. After removing all
6
Sanchis-Ojeda et al. 2015
FIG. 5.— In depth analysis of individual K2 transits. Upper left panel: Residuals of the K2 transit times with respect to a linear ephemeris, where the error bars
reflect only the formal uncertainties. The best solution has a final χ2 of 480 for 60 transit times. Our simulations show that this excess of scatter comes from
systematic uncertainties induced by the poor time resolution compared to the duration of the K2 transits. Lower left: The individual transit depths with time;
these confirm the erratic variations in the transit depths. Upper right: Scatter plot showing that there is no significant correlation between the timing residuals and
the transit depths. Lower right: the distribution of the 60 well-measured transit depths. This distribution is biased towards deeper transits, as transits shallower
than 0.2% were not analyzed due to the low S/N.
sample is added. The simulated datasets are then processed
with the same pipeline used as was used in this work for the
K2 data.
The main conclusions from these simulations are:
(i)
The formal uncertainties in the ‘measured’ transit depths
(∼0.05%) are indeed underestimates of the actual uncertain-
ties. The simulations using a constant depth have recovered
depths with a scatter of 0.1%. This is likely the result of a
combination of systematic effects induced by the short dura-
tion of the transits and the white noise terms. The simulations
show that an actual scatter in the depths greater than 0.15%
is easily recoverable. The scatter of our K2 depths is 0.2%,
so we conclude that it is real. The ground based observations
confirm this. (ii) The formal uncertainties also underestimate
the true uncertainties of the transit times (defined as the scat-
ter of measured times after removing a best linear trend). This
effect is worse in the presence of starspots, but it does not de-
pend on the transit depth variations. The effect can be large
enough to explain all the scatter observed in the K2 timing
analysis and therefore this scatter is not significant. We have
added a systematic uncertainty in quadrature to the timings of
3 minutes, chosen to provide a best-fit reduced χ2 of 1. (iii)
Our treatment of stellar spots does not induce detectable depth
variations (see Kawahara et al. 2013; Croll et al. 2015), as ex-
pected, mostly because we do not have the precision to detect
them.
3.2. Deviation from a standard transit profile
A fold of the activity corrected data (see section 2.3) about
the period we determined of 9.145704 hours is presented in
Fig. 6. The overall crudely triangular shaped transit profile
is the result of a convolution of the intrinsic shape and the
LC sampling time (see also Sects. 4 and 7). The depth of the
folded profile is 0.6%, and of course, this represents an aver-
age of the highly variable depths. The red curve is the mean
out-of-transit normalized flux (averaged for orbital phases be-
tween - 0.5 to - 0.25, and from 0.25 to 0.5). Note the clear
positive “bump” in flux just after the transit egress and the
smaller, but still marginally significant, bump just prior to
ingress. These are significant at the 6-σ and 2-σ confidence
limits, respectively.
These “bumps”, which are not normal features of exoplanet
transits, will be important for understanding the basic nature
of the transits (see Sect. 9). We call these features the “pre-
ingress bump” and the “post-egress bump”. Even in the cases
of KIC 1255b and KOI 2700b, the other two exoplanets which
appear to have dusty tails, the main distinguishing feature of
their transit profiles is a post-transit depression (Rappaport
et al. 2012; Rappaport et al. 2014) which is attributed to a
trailing comet-like dust tail. In addition, KIC 1255b exhibits
a pre-ingress “bump” which has been attributed to forward
scattering in the dust tail near the head of the dust cloud (see,
e.g., Rappaport et al. 2012; Brogi et al. 2012; Budaj 2013;
van Werkhoven et al. 2014). By contrast, the transit profile of
K2-22b
7
K2-22b has no post-egress depression, and the most promi-
nent bump comes after the egress, rather than before. These
features will be crucial to the interpretation of the dust “tail”
in this system.
impact parameter of b = 0.68 ± 0.06 (see Table 4) and a mean
depth of (R′
p can be understood as
the mean effective radius of the dust grains.
p/R∗)2 = 0.55%, where R′
3.3. Verification With Independently Processed K2 Data
We have also used the Vanderburg & Johnson (2014) flux
time series data set for K2-22 to check against the results of
our own pipeline. With the application of a simple high-pass
filter, this independently processed light curve looks nearly
identical to the one shown in the upper panel of Fig. 1. A
fold of the data about the period we have determined yields
a transit profile that is essentially the same as that shown in
Fig. 6, including the appearance of a convincing post-egress
bump, and a somewhat less significant pre-eclipse bump. To
the extent that our pipeline and that of Vanderburg & John-
son (2014) are independent, this is a satisfying test that our
processing has introduced no artifacts into the lightcurve.
4. GROUND-BASED TRANSIT OBSERVATIONS
A total of 12 transits were obtained from the ground using
1-m class telescopes, and the observations are summarized
in Table 1. The first observations were taken with the 1.2-
meter telescope at the Fred Lawrence Whipple Observatory
(FLWO) on Mt. Hopkins (AZ) using the KeplerCam instru-
ment, which has a single 4K × 4K Fairchild 486 CCD with
a 23′.1 × 23′.1 field of view. We successfully observed three
different transits with photometry in the Sloan i’ filter.
We also conducted a follow-up transit observation with
the Okayama 1.88-m telescope using the NIR camera ISLE
(Yanagisawa et al. 2006), and adopting a similar observing
scheme to that described in Fukui et al. 2014, we studied the
transit of the target in J-band. We only slightly defocused the
stellar images, but due to the faintness of the target and refer-
ence star, the raw counts were well within the detector’s linear
range. The typical FWHM of the target star’s PSF was ∼ 14
pixels, which corresponds to ∼ 3.5′′.
We obtained two additional transits with the IAC-80 (80-
cm) telescope at the Observatorio de Izaôsa, in the Canary
Islands. We used the wide field CAMELOT camera, with a
FOV of 10.4′ × 10.4′, with observations taken in the i-band.
No defocussing was applied. Weather conditions were clear
and stable through the two nights.
A total of six transits were observed with the 0.6-m TRAP-
PIST robotic telescope (TRAnsiting Planets and PlanetesI-
mals Small Telescope), located at ESO La Silla Observatory
(Chile). TRAPPIST is equipped with a thermoelectrically-
cooled 2K × 2K CCD, which has a pixel scale of 0.65′′ that
translates into a 22′ × 22′ field of view. For details of TRAP-
PIST, see Gillon et al. (2011) and Jehin et al. (2011). The ob-
servations were obtained through a blue-blocking filter26 that
has a transmittance of > 90% from 500 nm to beyond 1000
nm, without any additional defocus due to the faintness of the
target star. The procedures for the observation and data reduc-
tion are similar to those described by Gillon et al. (2013) and
we refer to this paper for further details.
Finally, additional observations were taken in queue mode
with OSIRIS@GTC on three different nights (see Table 1),
covering three complete transits of K2-22b. A spectroscopic
time series was taken in staring mode, starting ≈ 1 h before
the ingress, and finishing ≈ 1 h after the egress. The observing
logs are summarized in Table 1. Even though the GTC data
26 http://www.astrodon.com/products/filters/exoplanet/
FIG. 6.— Folded light curve of K2-22b for an orbital period of 9.14570
hours. The folded data have been averaged into 3.7-minute bins. The em-
pirical out of transit rms variations in the flux are 195 ppm. As explained in
the text there is a positive “bump” in flux just after the transit egress and a
smaller, but still marginally significant, bump just prior to ingress. The red
horizontal lines represent the mean out-of-transit normalized flux. The black
horizontal bar indicates the LC time of 29.4 minutes, and gives an indication
of the inherent temporal resolution of the light curve
Finally, even though we believe the transit is due to a dust
tail, as a baseline reference model we attempted to fit a stan-
dard transit profile of a solid planet over a limb-darkened star
(Mandel & Agol 2002) to our folded light curve. The high
distortion of the transit light curve due to the 30 minute sam-
pling precludes obtaining precise transit parameters, but we
were able to constrain the scaled semi-major axis from the
transit itself to be d/R∗ = 4.2+0.15
- 0.5 , with the uncertainties es-
timated from an MCMC analysis.
In fact, from the stellar
properties obetained in section 5.2, and Kepler’s 3rd law, we
can make a better direct estimate of d/R∗ = 3.3 ± 0.2, which
is compatible with the transit fit. We ran a final transit model
with a Gaussian prior on d/R∗ based on the inferred stellar
density, with mean value of 3.3 and a standard deviation of
0.2. This, in turn, allowed us to estimate a mean K2 total
transit duration (first to fourth contact) of 46 ± 1 minutes, an
8
Sanchis-Ojeda et al. 2015
were taken in low-resolution spectral mode, for the purpose
of transit timing, we used the data integrated over wavelength
(i.e., an effective white light transit). The spectral dependence
of the transits is discussed in detail in Sect. 7.
In all, 15 transits were measured from the ground. A log of
these 15 observations is given in Table 1.
Each of the above instruments has a different procedure for
reducing the light curves, and we refer to the corresponding
literature for a more detailed explanation (Okayama: Fukui et
al. 2011; TRAPPIST: Gillon et al. 2013; FLWO: Holman et
al. 2006; IAC80: Lázaro et al. 2015; GTC: see Sec. 7). All
transit light curves were obtained by comparing the fluxes of
the host star to a reference light curve made by combining up
to several comparison stars. They were all adjusted for dif-
ferential airmass corrections and a second-order polynomial
was fitted to the out-of-transit part of the light curves to re-
move long-term trends. The times of observation for all light
curves were transformed into BJD (TDB format, see Eastman
et al. 2010) to compare them with the K2 observations.
The 15 transit profiles measured from the ground are pre-
sented in Fig. 7. The profiles were fit with the same simple
three-segment transit model used in the K2 data analysis. The
parameters were transit depth, time of transit, transit dura-
tion, and ingress time. We also added two parameters to fit
for any fiducial linear trends with time.
In several cases a
transit fit barely represents an improvement over a straight-
line fit, whereas in some cases a deep transit is detected, con-
firming the depth variations (see Fig. 7). We first scaled the
uncertainties in the flux measurements to be equal to the stan-
dard deviation of the flux residuals with respect to the best
fit model. Correlated noise was taken into account using the
time-averaging method, in which the ratio of the standard de-
viation of the time-averaged residuals and the standard devia-
tion expected assuming white noise is calculated over a range
of timescales (Pont et al. 2006; Winn et al. 2008). In our
case, we obtained the final values of this β parameter as the
mean of the ratios with timescales from 10 to 30 minutes (see
Table 1), and multiplied the initial uncertainty by β when es-
timating the standard χ2 function. In those 9 cases where the
transit fit improves the minimum χ2 by at least 35, an MCMC
routine was used to estimate the uncertainties in the transit pa-
rameters. The 9 new transit times obtained from the ground
based observations are summarized in Table 2. We combined
the K2 transit times with the new 9 transit times to obtain the
final orbital ephemeris. The new transit times also show a
high level of scatter (χ2 = 30 for 9 O-C determinations).
The O-C values for all the well measured K2 and ground-
based transits are summarized in Fig. 8 (left panel). We used
these data to determine a best fit orbital period and its uncer-
tainty (after multiplying by the square root of the reduced χ2).
We repeated the process of fitting the O-C points, but this time
with a quadratic function, to set upper bounds on the deriva-
tive of the orbital period. The period and period derivative
results are summarized in Table 4.
The extra source of scatter in the ground-based transit tim-
ings could be caused by changes in the shape of the transit
light curve (see Croll et al. 2015). In principle these tim-
ing variations should be accompanied by transit duration or
shape variations, but our light curves are not precise enough
to allow the detection of such correlated variations (see right
panel of Fig. 8). We cannot discard the possibility that part
of the scatter is due to the use of a symmetric transit profile
to compute transit times, when indeed some of the transits
appear slightly asymmetric. From this set of 9 high-quality
ground-based transit measurements we obtained a weighted
total transit duration average of 50 ± 2 min, slightly longer
than the duration obtained from the K2 photometry. This dif-
ference of 4 ± 2 minutes, can be explained by the non-zero ca-
dence of the ground-based observations, which has not been
taken into account in these fits. A final value of 48 ±3 minutes
is quoted for the total transit duration in Table 4.
5. PROPERTIES OF THE HOST STAR
5.1. Imaging and color information
An early inspection of the SDSS images of the target star,
K2-22, showed that it is quite cool and likely an M star. The
image also indicated the presence of a faint companion at a
distance ∼2′′ in the South-West direction.
We observed the target star and its companion with Hyper-
Suprime Cam (HSC, Miyazaki et al. 2012) on the Subaru 8.2-
m telescope on 25 January, 2015 (UT). The sky condition on
that night was clear and photometric. The HSC is equipped
with 116 fully-depleted-type 2048 × 4096 CCDs with a pixel
scale of 0.17′′. We took images of 3 s and 60 s exposures
through g-, i-, and z-band filters (see Fig. 9). The host star
is saturated in i- and z-band with 60-s exposures, while the
companion star is not clearly seen in the g-band image with 3-
s exposure. We thus use the 60-s exposure images for g-band
and 3-s exposure images for i- and z-band for the analysis
presented here.
The presence of a stellar companion can be problematic,
particularly when the distance between the stars is smaller
than the pixel size of Kepler. We measure the flux ratios of
the companion to the host star in the HSC g-, i-, and z-band
images (see Fig. 9) by using the DoPHOT program (Schechter
et al. 1993), which performs PSF-fitting photometry while de-
blending stars. The background star is situated at a distance of
1.91′′, with a position angle of 134 degrees west from north
(see Table 3 for the coordinates). We obtain flux ratios of
4.0 ± 0.5%, 8.6 ± 0.5%, and 11.7 ± 0.5% for the g, i, and
z bands, respectively, where the uncertainties have been in-
creased to take into account systematic sources of noise.
The contribution of the background star points toward a
cooler and fainter companion. The bright star has been char-
acterized as part of the K2-TESS catalog, in which effective
temperatures are obtained from the colors of the stars (Stassun
et al. 2014). The temperature reported there for the brighter
target star is ∼ 3700 K, which means that both stars are likely
to be dwarfs stars (see Sect. 5.2). The colors used in that anal-
ysis could have been blended due to the proximity of the com-
panion. We also obtained the ugriz SDSS magnitudes and the
J 2MASS magnitude for both stars, confirming that the col-
ors used in Stassun et al. (2014) were correct. We re-derived
the flux ratios in g, i and z band (5.4%, 6.5% and 12% re-
spectively) confirming the measurements obtained using the
Subaru images. The minimum uncertainties used in the fits
are 0.05 magnitudes to take into account systematic effects.
Adopting the full line-of-sight extinction of AV = 0.17 based
on the expected distance to the objects, and a log g of 5 due
to their red colors and dwarf nature (see Sect. 5.2), we ob-
tained temperatures of 3800 ± 150 K for the bright star and
3150 ± 200 K for its fainter companion.
The magnitudes and colors of the two stars based on a com-
bination of the SDSS, Subaru, and 2-MASS photometry are
summarized in Table 3.
K2-22b
9
GROUND BASED TRANSIT OBSERVATIONS OF K2-22B
TABLE 1
Date
[UT]
Start
End
Epoch
Telescope
Instrument
or filter
Number of
data points
Median time
between points [min]
Airmassa
σb
[ppm]
βb
Sloan i’
J-band
Sloan i’
Sloan i
R1000R
07:30
14:40
07:13
02:33
03:15
03:24
02:16
02:32
00:10
03:24
05:22
03:27
02:27
04:38
02:00
12:45
18:23
12:06
05:55
06:45
09:35
07:06
06:28
03:30
06:42
12:30
07:39
05:24
08:48
05:37
2015 Jan 15
2015 Jan 17
2015 Jan 23
2015 Jan 27
2015 Jan 29
2015 Feb 02
2015 Feb 04
2015 Feb 12
2015 Feb 14
2015 Feb 15
2015 Feb 18
2015 Feb 23
2015 Feb 25
1
2015 Feb 26
1
2015 Mar 21
1
a The airmass range is shown as z0 → zmin → zfin, where z0 and zfin represent the airmass at the beginning and at the end of the night, respectively, and zmin
represents the minimum airmass.
b σ refers to the flux scatter respect to the best fit model, whereas β represents the level of correlated noise (see section 4).
1.89 → 1.15 → 1.25
2.32 → 1.18 → 1.69
1.78 → 1.15 → 1.24
1.22 → 1.13 → 1.28
1.12 → 1.11 → 1.53
2.34 → 1.18 → 1.42
1.18 → 1.11 → 1.79
2.60 → 1.18 → 1.18
1.45 → 1.11 → 1.12
1.12 → 1.12 → 2.07
1.88 → 1.15 → 1.90
1.50 → 1.18 → 1.30
1.71 → 1.18 → 1.18
1.23 → 1.18 → 1.69
1.43 → 1.18 → 1.26
1.2 m FLWO
Okayama
1.2 m FLWO
IAC-80
GTC
TRAPPIST
IAC-80
TRAPPIST
GTC
GTC
1.2 m FLWO
TRAPPIST
TRAPPIST
TRAPPIST
TRAPPIST
2800
3000
3500
4070
570
5500
4300
4070
920
690
2270
2600
3230
2760
2800
Blue blocking
Blue blocking
Blue blocking
Blue blocking
136
102
126
117
44
314
260
207
40
38
188
213
137
202
182
595
601
616
626
634
642
647
668
673
676
684
697
702
705
765
2.3
2.1
2.3
1.7
4.6
1.1
1.1
1.1
5.4
5.4
2.3
1.1
1.1
1.1
1.1
1
1
1
1.24
1.13
1.15
1.00
1.52
1.46
1.26
1.50
1.38
Blue blocking
Sloan i
Blue blocking
R1000R
R1000R
Sloan i’
TABLE 2
GROUND BASED TRANSIT TIMES
Epoch
Time of transit (BJD) Uncertainty (days)
595
626
634
647
673
676
684
697
705
2457037.8608
2457049.6705
2457052.7269
2457057.6831
2457067.5839
2457068.7295
2457071.7778
2457076.7327
2457079.7805
See section 4 for details.
0.0007
0.0022
0.0008
0.0021
0.0015
0.0005
0.0018
0.0012
0.0012
PROPERTIES OF THE HOST STAR AND COMPANION
TABLE 3
Parameter (units)
Host star
Companion
11:17:55.763
+02:37:05.48
21.68 ± 0.17
19.61 ± 0.05
18.79 ± 0.05
17.34 ± 0.05
16.34 ± 0.05
14.87 ± 0.05
14.27 ± 0.05
13.93 ± 0.05
3290 ± 120
11:17:55.856
+02:37:06.79
19.07 ± 0.05
16.44 ± 0.05
15.01 ± 0.05
14.38 ± 0.05
14.05 ± 0.05
12.74 ± 0.05
12.09 ± 0.05
11.91 ± 0.05
3830 ± 100
4.65 ± 0.12
0.03 ± 0.08
0.60 ± 0.07
0.57 ± 0.06
0.063+0.008
- 0.007
M0V ±1
225 ± 50
RA (J2000)
DEC (J2000)
u-mag (SDSS)
g-mag (SDSS)
r-mag (SDSS)
i-mag (SDSS)
z-mag (SDSS)
J (2MASS)
H (2MASS)
KS (2MASS)
Teff (K)
log g
[Fe/H]
M∗ (R⊙)
R∗ (M⊙)
L∗ (L⊙)
Spec Type
Distance (pc)
The magnitudes and colors are taken from a combination
of the SDSS and 2-MASS photometry. The stellar parame-
ters are inferred from the combined analyses of the Keck-
HIRES, IRTF-SpeX, and UH88-SNIFS spectra. See sec-
tion 5 for details.
0.06 ± 0.20
0.27 ± 0.05
0.30 ± 0.08
0.010+0.007
- 0.005
M4V ±1
225 ± 50
...
TABLE 4
PROPERTIES OF K2-22B
b
Parameter
Orbital Perioda (days)
Orbital Period (hr)
Transit centera (BJD)
Porb/Porb (yr- 1)a
Total transit durationa (min)
d/R∗
Impact parameter, bc
d (AU)
Rp (R⊕)d
Mp (MJ)e
θ∗
Mdust (g s- 1)g
ℓ2 (leading tail only; units of R∗)h
agrain (µm)h
Impact parameter, bh
ρ1/ρ2 (two-tail model, ℓ1 = ℓ2)h
f
Value
0.381078 ± 0.000001
9.145872 ± 0.000024
2456811.1208 ± 0.0006
. 3.5 × 10- 7
48 ± 3
3.3 ± 0.2
0.68 ± 0.06
0.0088 ± 0.0008
< 2.5 ± 0.4
< 1.4
17◦ ± 1.5◦
≈ 2 × 1011
0.19 - 0.48
0.3 - 0.5
0.42 - 0.78
< 0.5
a Derived from the K2 and ground-based observations.
b Based on the mass and radius of the host star given in Table 3
and Kepler’s 3rd law.
c Derived from the K2 observations using a standard Mandel &
Agol (2002) fit to a hard-body transit.
d Based on the shallowest K2 transits.
e 2-σ limit based on the Keck RVs.
f Estimate of the half angle subtended by the star at the position
of the planet.
g Following the type of estimate made in Appendix D of Rap-
paport et al. (2014)
h 90% confidence limits based on the models of section 9. The
leading and trailing tails are assumed to have exponential scale
lengths and maximum optical thicknesses, ℓ2, ℓ1, ρ2, and ρ1,
respectively.
5.2. Spectral Studies
5.2.1. NOT-FIES spectrum
An exploratory spectrum was obtained on 13 Feb 2015
with the FIbre-fed Échelle Spectrograph (Frandsen & Lind-
berg 1999, Telting et al. 2014) mounted at the 2.56-m Nordic
Optical Telescope of Roque de los Muchachos Observatory
(La Palma, Spain). We used the 1.3′′ Med-Res fibre which
provides a resolving power of R = 46000 over the spectral
10
Sanchis-Ojeda et al. 2015
FIG. 7.— Fifteen transits of K2-22b observed from the ground, plotted to the same vertical scale. The transit models are each vertically offset by 0.02 for
clarity in presentation. When necessary, the observations are binned to have a cadence close to 5 minutes. In spite of the weaker statistics for some of these, it is
apparent that the transit depths vary considerably.
FIG. 8.— Left panel: Timing residuals after removing the best fit linear orbital ephemeris. The uncertainties on the K2 timings have been increased to take into
account systematic effects. Right panel: The uncertainties in the transit durations are too high to allow for a detection of a possible correlation between transit
duration and transit time.
K2-22b
11
ination from the calibration lamp.
The photospheric parameters of the template stars were ho-
mogeneously derived from the SOPHIE spectra using the pro-
cedure described in Maldonado et al. (2015), which relies
on the ratios of pseudo-equivalent widths of different spec-
tral features. Unfortunately, the low S/N ratio prevented us
from directly applying this technique to the co-added HIRES
spectrum of the target.
Prior to the fitting procedure, the resolution of the tem-
plate spectra was somewhat degraded to match that of the
HIRES spectrograph (R = 60 000) – by convolving the SO-
PHIE spectra with a Gaussian function mimicking the dif-
ference between the two instrument profiles. A corrective
radial velocity shift was estimated by cross-correlating the
observed and template spectra. We restricted the spectral
range over which the fit is performed to 5500–6800 Å and
masked out the regions containing telluric lines. We se-
lected the 5 best fitting templates and adopted the weighted
means of their spectroscopic parameters as the final esti-
mates for the target star. We found that the target star has
an effective temperature of Teff = 3780 ± 90 K, surface grav-
ity of log g = 4.65 ± 0.12 (log10 cm s- 2), and iron abundance
of [Fe/H] = 0.05 ± 0.08 dex. We also set an upper limit of
1.5 km s- 1 on the projected rotation velocity v sin i⋆ by fitting
the profile of several clean and unblended metal lines to the
PHOENIX model spectrum (Husser et al. 2013) with the same
parameters as the target star. Figure 10 shows the co-added
HIRES spectrum in the spectal region around the Hα line,
along with the best fitting SOPHIE template.
FIG. 9.— Subaru/HSC+z image of K2-22 shows a secondary companion
star ∼ 2′′ away, with a position angle of 134 degrees west from north. In this
band, the flux ratio between the faint and bright star is 12%.
range 3640–7360 Å. The FIES data revealed a single-lined
spectrum. Although the low signal-to-noise ratio does not al-
low us to perform a reliable spectral analysis, a comparison
with a grid of stellar templates from (Valdes et al. 2004) and
(Bochanski et al. 2007) confirmed that the host is a cold dwarf
star.
5.2.2. Keck-HIRES spectra
5.2.3. IRTF-SpeX spectrum
We also acquired five spectra of the host star with the Keck
Telescope and HIRES spectrometer using the standard setup
of the California Planet Search (CPS, Howard et al. 2010).
Over the course of four nights, (5-8 Feb 2015) we observed
under clear skies and average seeing (∼1.2′′). Exposure times
of less than 5 minutes resulted in SNRs of 5-10 per pixel.
Extensive scattered light and sky emission were unavoidable
for exposures on 5 Feb 2015 due to the close proximity of the
nearly full moon to the target star.All spectra were taken using
the C2 decker, which is 0.87′′ wide by 14′′ long, resulting in
a spectral resolution of R=60,000. The slit was oriented to
minimize the amount of contamination from the companion
star.
We estimated the effective temperature Teff, surface gravity
log g, iron abundance [Fe/H], and projected rotation veloc-
ity v sin i⋆ of the host star from the co-added HIRES spec-
trum, which has a S/N ratio of about 20 per pixel at 6000 Å.
We used a modified version of the spectral analysis tech-
nique described in Gandolfi et al. (2008), which is based on
the use of stellar templates to simultaneously derive spectral
type, luminosity class, and interstellar reddening from flux-
calibrated, low-resolution spectra. We modified the code to
fit the co-added HIRES spectrum to a grid of templates of
M dwarfs recorded with the SOPHIE spectrograph (Bouchy
et al. 2008). We retrieved from the SOPHIE archive27 the
high-resolution (R = 75 000), high S/N ratio (>80) spectra of
about 50 bright red dwarfs encompassing the spectral range
K5–M3 V. The stars were selected from the compilation of
(Lepine et al. 2013). We downloaded spectra with no simulta-
neous thorium-argon observations, to avoid potential contam-
27 Available at http://atlas.obs-hp.fr/sophie/
A near-infrared spectrum of the star was also obtained using
the updated SpeX (uSpeX) spectrograph (Rayner et al. 2003)
on the NASA Infrared Telescope Facility (IRTF). SpeX ob-
servations were taken using the short cross-dispersed mode
and the 0.3′′ × 15′′ slit, which provides simultaneous cover-
age from 0.7 to 2.5 µm at R ≃ 2000. The slit was aligned to
capture both the target and companion spectrum. The pair
was nodded between two positions along the slit to subse-
quently subtract the sky background. Ten spectra were taken
following this pattern, which provided a final S/N of ≃ 100
per resolving element in the K-band for the primary, and ≃
25 for the companion. The spectra were flat fielded, extracted,
wavelength calibrated, and stacked using the SpeXTool pack-
age (Cushing et al. 2004). An A0V-type star was observed
immediately after the target, which was used to create a tel-
luric correction using the xtellcor package (Vacca et al. 2003).
We analyzed the SpeX spectra obtained for both the bright
and faint star, and both show strong atomic and weak CO
absorption, as is expected for dwarf stars. Comparison with
dwarf and giant NIR templates from the IRTF library (Rayner
et al. 2009) rule out the possibility of either component being
evolved. Metallicity was derived from the SpeX data using the
procedures from Mann et al (2013), who provide empirical re-
lations between atomic features and M dwarf metallicity, cal-
ibrated using wide binaries. Teff was calculated using the em-
pirical calibration from Mann et al. (2013), which is based on
stars with Teff determined from long-baseline optical interfer-
ometry (Boyajian et al. 2012). This analysis yielded a metal-
licity of 0.00±0.08 and a Teff of 3880±85 K for the primary;
and a metallicity of 0.06±0.20 and a Teff of 3290±120 K for
the companion. Both the metallicity and Teff determinations
for the primary star are consistent with those derived from the
12
Sanchis-Ojeda et al. 2015
spread in rotation periods of stars in clusters (e.g., Reiners
& Mohanty 2012). However, the observed rotation period of
the target star, 15.3 days (see Sect. 2.3), sits slightly above
the main period distribution of M dwarfs in the Hyades and
Praesepe clusters (Delorme et al. 2011), possibly indicating
an older age. On the other hand, McQuillan et al. (2013)
finds a bimodal distribution of the rotation periods of field
M dwarfs, likely associated with two populations with differ-
ent median ages. K2-22 belongs to the shorter period (10-25
days), younger stellar population group. Both observations
suggest that the age of the system could be between 1 and a
few Gyrs.
The overall best-determined physical properties of the two
stars, based on the spectroscopic observations and analyses,
including M∗, R∗, L∗, Teff, log g, metallicity, and distance, are
given in Table 3.
6. RADIAL VELOCITIES
The five Keck HIRES spectra obtained can also be used
to place constraints on the mass of the putative planet. The
standard CPS pipeline is used to convert from raw spectra to
two-dimensional spectra. Each of the three HIRES CCDs are
independently reduced with flat fields, sky subtraction, and
cosmic-ray removal. The pixel columns at each wavelength
are then summed, resulting in flux as a function of wavelength
for each pixel. Consistent wavelength solutions are insured by
aligning a carefully chosen set of Thorium-Argon emission
lines onto the same pixels at the beginning of each night’s
observations.
The systemic radial velocity of each star is measured using
the A-band and B-band telluric line features. Using the tel-
luric lines as the wavelength fiducial, the relative placement
of the stellar absorption lines is measured, and referenced to
stars of known radial velocity (Chubak et al. 2012). These
radial velocity measurements are made relative the Earth’s
barycenter and are accurate to ±0.3 km/s. No RV variabil-
ity in phase with the orbit of the target is detected, and the
radial velocities have an rms of 0.3 km/s, compatible with the
expected uncertainties.
We used these five Keck HIRES points to set a formal 1-σ
upper limit on the RV amplitude of the host star of 280 m/sec,
yielding a 2 - σ upper limit on the planet mass of 1.4 MJ.
This is not highly constraining in the context of a small rocky
planet, but it does rule out non-planetary scenarios (assuming
that the source of the photometric dips is the brighter star).
7. GTC MULTICOLOR OBSERVATIONS
7.1. Wavelength Dependent Transits
Spectro-photometric observations for 3 complete transits
were obtained with OSIRIS on the GTC (see also Sect. 4).
The GTC instrument OSIRIS consists of two CCD detectors
with a field of view (FOV) of 7.8′ × 7.8′ and a plate scale of
0.127′′ per pixel. For our observations, we used the 2 × 2
binning mode, a readout speed of 200 kHz with a gain of 0.95
e-/ADU and a readout noise of 4.5 e-. We used OSIRIS in
its long-slit spectroscopic mode, selecting the grism R1000R
which covers the spectral range of 520-1040 nm with a res-
olution of R = 1122 at 751 nm. The observations and results
we present here were taken using a custom built slit of 12′′
in width, with the target and a comparison star both located
in the slit. The use of a wider slit has the advantage of re-
ducing the possible systematic effects that can be introduced
by light losses due to changes in seeing and/or imperfect tele-
scope tracking (Murgas et al. 2014).
FIG. 10.— HIRES co-added spectrum of K2-22 (black line) encompassing
the Hα line. The best fitting template spectrun is overplotted with a thick red
line.
analysis of the HIRES spectra.
5.2.4. UH88 SNIFS spectra
Finally, spectra of both K2-22 and its companion star were
obtained with the SNIFS integral field spectrograph on the
UH88 telescope on Mauna Kea during the night of UT 31
March 2015. The two stars were spatially resolved in the im-
age cubes. SNIFS spectra cover 3200-9700 Å with R ≈ 1000,
do not suffer from slit effects, and have been precisely cali-
brated by extensive observations of spectrophotometric stan-
dards (Lantz et al. 2004; Mann et al. 2011; Mann et al. 2013).
The wavelength coverage and resolution of SNIFS is more
than adequate for measuring the strength of key molecular
bands and atomic lines as indicators of Teff and gravity for
M dwarf stars. SNR > 100 was obtained for the primary.
The effective temperature of the target star was derived in-
dependently from the SNIFS spectrum by comparing it to
Dartmouth Stellar Evolution model predictions in a man-
ner that has been calibrated to retrieve the bolometrically-
determined temperatures of nearby stars with measured an-
gular radii (Boyajian et al. 2012; Mann et al. 2013). Radius
and mass were then derived from empirical relations based
on an expanded set of calibrator stars (Mann et al. 2015),
with masses obtained from the Delfosse et al. (2000) sam-
ple. These yield Teff = 3780 ± 60 K, R∗ = 0.55 ± 0.03 R⊙,
M∗ = 0.57 ± 0.06 M⊙, and log g = 4.72 ± 0.05, in quite good
agreement with the other determinations of these parameters
discussed above.
Adopting the mean [Fe/H] and Teff derived from SpeX and
HIRES we derived R∗, M∗, and L∗ for the primary star, and we
did the same with the SpeX parameters for the secondary star.
To this end we utilized empirical Teff-[Fe/H]-R∗, Teff-[Fe/H]-
M∗, and Teff-[Fe/H]-L∗ relations derived from the Mann et
al. (2015) sample. Accounting for errors in Teff, [Fe/H] and
the Mann et al. (2015) relations we computed physical pa-
rameters that are listed in Table 3.
We also derived independent estimates of the distance to the
stars, based on the colors and also on the spectroscopic param-
eters, reaching similar values with different methods and for
both stars. We conclude that both stars are likely bound and
at a distance of 225 ± 50 pc.
Constraints on the age of the system can be derived from
the stellar rotation period using gyrochronology (see, e.g.,
Barnes 2007). It has been suggested that the classical rela-
tionships may not work for M dwarfs because of the large
K2-22b
13
During the first transit observed with the GTC (29 Jan) we
took as a reference a very close comparison star to the east of
the target. However, for the two last transits (13 and 14 Feb)
we took a reference star with a similar brightness to K2-22
and located at a distance of 2.7′ from the target. The posi-
tion angle of the reference star with respect to the target was
- 79.2◦. The two stars were positioned equidistantly from the
optical axis, close to the center of CCD#1, while CCD#2 was
turned off to avoid crosstalk.
The basic data reduction of the GTC transits was per-
formed using standard procedures. The bias and flat field
images were produced using the Image Reduction and Anal-
ysis Facility (IRAF28) and were used to correct the images
before the extraction of the spectra. The extraction and wave-
length calibration was made using a PyRAF29 script written
for GTC@OSIRIS long-slit data. This script automated some
of the steps to produce the spectra such as: extraction of each
spectrum, extraction of the corresponding calibration arc, and
wavelength calibration (using the HgAr, Xe, Ne lamps pro-
vided for the observations). All spectra were aligned to the
first spectrum of the series to correct for possible shifts in the
pixel/wavelength solution during the observations caused by
flexures of the instrument. Several apertures were tested dur-
ing the reduction process, and the one that delivered the best
results in terms of low scatter (measured in rms) in the points
outside of transit for the white light curve was selected. The
results presented here were obtained using apertures of 28,
40, and 44 pixels in width for the three transit observations,
respectively. Final spectra were not corrected for instrumen-
tal response nor were they flux calibrated. Fig. 11 shows the
extracted spectrum of K2-22.
FIG. 11.— Visual explanation of our different choices for splitting the GTC
observations into several different broad color bands. These are centered at
652.5 nm and 820 nm for the ‘red’ and ‘blue’ colors, and 625 nm, 735 nm,
and 850 nm respectively, for the 3-color bands.
The universal time of data acquisition was obtained using
the recorded headers of the spectra indicating the opening and
closing time of the shutter in order to compute the time of
mid exposure. We then used the code written by Eastman et
28 IRAF is distributed by the National Optical Astronomy Observatory,
which is operated by the Association of Universities for Research in Astron-
omy (AURA) under a cooperative agreement with the National Science Foun-
dation.
29 Python environment for IRAF
MULTICOLOR OBSERVATIONS SUMMARY
TABLE 5
Date
δ570- 680nm
δ680- 790nm
δ790- 900nm
α
Jan 29
0.50 ± 0.06% 0.47 ± 0.05% 0.46 ± 0.05% 0.13 ± 0.55
Feb 14
0.48 ± 0.04% 0.49 ± 0.03% 0.44 ± 0.03% 0.11 ± 0.37
Feb 15
0.95 ± 0.04% 0.83 ± 0.03% 0.72 ± 0.04% 0.83 ± 0.23
These transit depths are measured from the data alone, and do not take into
account contamination from the faint companion or airmass corrections.
The calculated Angström exponent, α, is, however, corrected for the small
dilution factor of the faint companion star.
al. (2010)30 to compute the BJD time using the mid-exposure
time for each of the spectra to produce the light curves ana-
lyzed here.
We constructed light curves over several spectral ranges to
search for color dependencies of the transit shape and depth.
In the case of the white light curve, the flux was integrated
over almost the entire wavelength range of the observed spec-
tra, between 570 and 900 nm, but avoiding the blue and red
ends of the spectra where the SNR is lower. This step is
particularly important, because observations at redder wave-
lengths suffer from fringing31 whereas using the data at bluer
wavelengths not only did not improve the quality of the light
curves, but increased the occurrence of outliers. Three nar-
rower color light curves were created from the raw spectra for
each star, by integrating all counts over the spectral ranges
570-680 nm, 680-790 nm, and 790-900 nm (see Fig. 11).
The final light curves used in the analyses were created by
dividing the K2-22 light curves by the comparison star light
curves. The scatter (std) for the relative white light curves can
be found in Table 1.
The white light transit profiles of the three separate GTC
transits were shown earlier in Fig. 7. It is obvious that the
depth of the transits change from one night to the next, con-
firming the rapidly varying nature of this object. Note that
between transits 2 and 3, only 27.44 hours had passed and the
transit depth nearly doubled.
To further investigate the nature of these changes in the
GTC transit curves, in Fig. 12 we plot the three color light
curves for each transit separately. Transit 3 (bottom panel) is
the deepest and shows a clearly increasing transit depth to-
wards the blue. The SNR is lower in the other two sets of
color light curves and the transits are shallower, and thus there
is no clear trend in transit depth with wavelength. In order
to interpret quantitatively the color dependence of the depths
in the third GTC transit, we fit our standard three-segment
transit profile model to the data from each of the nights, with
three different transit depths to evaluate how the transit depth
changes with color. We also fit each transit with five addi-
tional parameters, to construct a polynomial that reaches sec-
ond order in time from mid-transit and it linearly depends on
the mean-subtracted airmass and full-width half maximum of
the images of the host star. The total number of parameters
is 21 for each observation, with a total 120, 132 and 114 data
points in each of the three nights. After finding the best-fit
model, we set the error of the flux measurements of each light
curve to provide a best-fit standard χ2 value equal to the num-
ber of degrees of freedom. We then took correlated noise into
account computing the β factors for each light curve and each
night (see section 4), and multiplying the errors by their cor-
30 http://astroutils.astronomy.ohio-state.edu/time/
31 http://www.gtc.iac.es/instruments/osiris/osiris.php#Fringing
14
Sanchis-Ojeda et al. 2015
FIG. 12.— GTC observations split among three color bands. On the left panels we show the raw light curves, which have been median-normalized. The
observations shown on the right hand side have been corrected for time, airmass and seeing effects. The first two transits were shallower, and depth variations are
hard to notice, but during the third night, when the white light transit is deepest, color variations are clearly observed.
responding β factor. We use an MCMC routine to obtain pos-
terior distributions for the transit depths, which are marginal-
ized respect to all the other 18 model parameters (see Table 5).
From the marginalized posterior distributions we compute
a quantity called the “Angström exponent”, α, which is de-
fined as - d ln σ/d ln λ , where σ is the effective extinction
cross section. Here we take the transit depth (on each night)
at wavelength, λ, to be proportional to the effective cross sec-
tion at λ (under the assumption that the dust tail is optically
thin). In order to use the transit depths, we first multiply each
individual transit depth by (1 + Di) to apply a dilution correc-
tion, where each Di is the flux ratio between the faint com-
panion and the host star evaluated at the flux weighted center
of each selected band. These Di values are obtained from our
SED models (see Sect. 5.1), and have values of 4.4%, 6.6%
and 8.9% with increasing wavelength, with uncertainties of
0.2%. We then find α = 0.83 ± 0.23, for the transit observed
on February 15. For the other two transits, the values of α
are 0.11 ± 0.37 and 0.13 ± 0.55, both slightly positive, but
not statistically significantly different from zero. We made
no corrections for the wavelength-dependent limb darkening
properties, but utilized the quadratic limb darkening coeffi-
cients of Claret & Bloemen (2011) to estimate that the val-
ues of α would be lowered by between 0.15 and 0.02, for an
equatorial transit (which is unlikely) and an impact parameter
of b = 0.7, respectively. For higher impact parameters, up to
0.85 (see Table 4), the value of α could actually be raised by
up to 0.15, and become even more significantly different from
zero.
7.2. Interpretation in Terms of Dust
We interpret the value of the Angström exponents in terms
of Mie scattering (for spherical dielectric dust particles) with a
variety of different compositions. We computed the Angström
exponent, α, over the range 630-840 nm for a set of power-law
distributions for the dust particle sizes, with dN/da ∝ a- Γ. In
order to guarantee that the total cross section converges, we
also need to specify a maximum grain size, amax. For the sake
of specificity we adopted an illustrative dust composition of
corundum, but the conclusions we draw are the same for a
number of other common refractory materials, and in fact,
for any material with real and imaginary indices of refrac-
tion of n ≃ 1.6 and 0.001 . k . 0.03. We plot the computed
Angström exponent in Fig. 13 as a function of amax for five
different power-law exponents, Γ. As can be seen from the
figure, values of α in the range of ∼0 - 1, as indicated in Ta-
ble 5, correspond to non-steep power-law indices of Γ ≃ 1 - 3
and maximum particle sizes of ∼0.4 - 0.7 µm. In turn, these
correspond to “effective particle sizes” of 0.2 µm to 0.4 µm,
where the effective radius is the average grain radius weighted
by both the size distribution and the cross section, i.e.,
aeff =Z amax
0
a1- Γ σ(a, λ) da /Z amax
0
a- Γ σ(a, λ) da
(1)
where σ(a, λ) is the wavelength dependent Mie extinction
cross section for a particle of radius a.
FIG. 13.— The Angström exponent computed from Mie scattering over
the wavelength range 0.63 - 0.84 µm as a function of the maximum particle
size, for five different indices, Γ, of the power-law grain size distribution, i.e.,
dN/da ∝ a- Γ. We adopted an illustrative particle material of corundum.
8. WHY THE BRIGHTER STAR IS THE SOURCE OF THE TRANSITS
Here we summarize why we are confident that it is the
brighter star that is the source of the observed transits. First,
the fainter star is redder than the bright star (see Sect. 5.1); in
K2-22b
15
particular the contribution from the faint star to the total flux
increases by a factor of 2 from r to i band and a factor of 4
from r to z band. If we assume that the faint star is the host,
and assume an achromatic mean transit depth, the changes in
flux ratio should translate into correspondingly larger transit
depths toward the red due to the dilution effect. In section 7
we showed simultaneous GTC transit observations in bands
similar to r and a combination of the i and z bands. The tran-
sit depth, at least on one occasion, is actually considerably
shallower in the latter redder band, strongly suggesting that
the bright star is the one being transited.
From another line of argument, we note that the transit
depths are sometimes as large as 1.3%, and if the fainter star
(5% of the flux in g-band) were the source of the transits, it
would have to be attenuated by 24% of its flux. Since the
transits are highly variable in depth, and the transit profile is
not that due to a conventional hard-body transit, it is almost
certainly due to a ‘soft’ attenuator such as a dust tail. But, the
fainter star has a radius of 0.3 R⊙, and therefore an occulting
dust tail would have to either cover the entire star with an op-
tical depth, τ, of ∼0.25, or cover ∼25% of the star (i.e., a tail
thickness of ∼0.08 R⊙) with τ ≃ 1. Simulations of the dust
tail in KIC 1255b (Rappaport et al. 2012) and this object (see
Sect. 9.2) indicate that the vertical thickness of the dust tail is
only ∼0.03 R⊙. This would make the scenario of a dust tail
covering over 1/4 of the area of the fainter star highly unlikely.
9. THE DISINTEGRATING PLANET HYPOTHESIS
9.1. Evidence for a Disintegrating Planet
When taken together, all these observational results point
clearly toward a planet in a 9-hour orbit that is disintegrat-
ing via the emission of dusty effluents. The lines of evi-
dence pointing in this direction include: (1) erratically and
highly variable transit depths (see Sect. 3); (2) transit profile
shapes from ground-based observations that are likely vari-
able (though with lesser confidence that the depth changes;
see Sects. 4 and 7); and (3) an average transit profile from
the K2 data that exhibits clear evidence for a post-transit
“bump” and also weaker evidence for a pre-transit “bump”
(Sect. 3.2). The first of these is highly reminiscent of the dis-
integrating planet KIC 1255b (Rappaport et al. 2012), while
the variable transit shapes are also detected in KIC 1255b
from ground-based studies (R. Alonso et al. private communi-
cation; Bochinski et al. 2015). The transit profile of K2-22b,
with a post-transit “bump”, is different from the transit pro-
files of KIC 1255b and KOI 2700b which show a post-transit
depression as opposed to a post-transit bump. The first two
of the above listed features point to obscuration by dusty ef-
fluents coming from a planet, while the third property needs
to be explained in this same context. In the following sec-
tions we explore the significance and the interpretation of the
transit profile.
9.2. Quantitative Model for the Leading Dust Tail
A dust tail emanating from a planet, as inferred in the cases
of KIC 1255b and KOI 2700b, trails the planet as is illustrated
in Fig. 6 of Rappaport et al. (2012; see also the middle panel
of our Fig. 15). Such dust tails are the way they would be seen
in the reference frame of the planet, and note that the motion
of the planet is implicitly in the opposite direction from the
tail. In this case we would say that the tail “trails the planet”.
The reason for this is that the radiation pressure acting on the
dust forces it into an eccentric orbit with its periastron located
at the point where the particle was released. This orbit has a
larger semimajor axis than that of the planet (see Appendix B
of Rappaport et al. 2014). In turn, the larger orbit has a lower
orbital frequency, and the particles appear to trail behind the
planet in a comet-like tail.
An observer viewing this system is effectively moving
counterclockwise. Thus, the ingress to the transit is sharp as
the “comet head” moves onto the stellar disk (see Fig. 6 in
Rappaport et al. 2012). The trailing tail would lead to a de-
pression upon egress as the tail slowly moves off of the stellar
disk. In that case, the pre-transit “bump” is caused by forward
scattering by the densest regions of the dust which have not
yet reached the stellar disk. This is the case we believe we see
in KIC 1255b (Rappaport et al. 2012; Brogi et al. 2012; Bu-
daj 2013; van Werkhoven et al. 2014). If the tail of the planet
is sufficiently short (i.e., compared to the radius of the host
star) then there would be both a pre-transit and a post-transit
“bump”, the latter of which would dominate over the rela-
tively shallow post-transit depression (see also Budaj 2013).
In this case, the pre-transit bump would be somewhat larger
due to the asymmetry in the direction of the tail.
A logical first guess as to how to produce a post-transit
“bump” on the transit curve would be to reverse the direction
of the comet-like tail. However, as we have seen, substan-
tial radiation pressure forces inevitably lead to a trailing tail.
Then, the question becomes how to produce a “leading dust
tail” to the planet. One way to have dusty material lead the
planet, i.e., moving faster, would be to have it overflow its
Roche lobe (or, Hill sphere of influence) and fall in toward
the host star. At first consideration this doesn’t seem to work
since a rocky planet in a 9-hour orbit will not be close to filling
its Roche lobe. Rappaport et al. (2013) showed that the criti-
cal density for Roche-lobe overflow is largely a function of its
orbital period, and is nearly independent of the properties of
the host star. In particular, Eqn. (5) in Rappaport et al. (2013)
suggests that critical density for a planet to be filling its Roche
lobe is
Porb (cid:19)2
ρcrit ≃(cid:18) 11.3 hr
g/cm3 ≃ 1.5 g/cm3
(2)
where the right-hand value is for a 9-hour planet. It seems
very unlikely that a planet with this low a mean density would
be disintegrating via dusty effluents. Turning the problem
around, we can ask what fraction of the Roche-lobe radius
would be occupied by a planet with a mean density in the
range of 5-8 g/cc, which might be more appropriate for a dust
emitter (see, e.g., Rappaport et al. 2013). It is then evident that
for densities which are ∼3 - 5 times higher than their critical
densities, their radii are not even factors of ∼2 times smaller
than their Roche lobes.
Therefore, we come to the conclusion that planets with
rocky compositions in a 9-hour orbit are underfilling their
Roche lobes by only a factor of ∼2. This is more than suffi-
cient to prevent the planet from directly overflowing its Roche
lobe. However, the potential difference between the planet’s
surface and the Roche lobe is about half the potential differ-
ence to infinity. Thus, if the mechanism that drives off the
dust or the heavy metal vapors that condense into dust, e.g.,
via a Parker-type wind (Rappaport et al. 2012; Perez-Becker
& Chiang 2013) imparts the full escape velocity of the planet
or more, then the material can certainly reach the surface of
its Roche lobe.
If the material reaching the Roche surface feels a substantial
16
Sanchis-Ojeda et al. 2015
FIG. 14.— Calculated values of β, the ratio of radiation pressure to grav-
itational forces, as a function of particle size. The blue and green curves
(nearly superposed) are for KIC 1255b and KOI 2700b, respectively. The
red curves are values of β for EPIC 201673175, for different materials. For
KIC 1255b and KOI 2700b we adopted n = 1.65 and k = 0.01 for the real and
imaginary parts of the index of refraction. The solid red curve is for the same
indices. The dot-dashed curve is for the same real index but with k = 0.02,
while the dotted, dashed, and dot-dot-dashed curves are specifically for iron,
corundum and forsterite, respectively. The horizontal black dotted line indi-
cates the value of β = 0.02 below which most dust particles would go into a
leading tail.
radiation pressure, it will still be blown back into a comet-like
tail. By “substantial” we mean that β, the ratio of radiation
pressure forces to gravitational forces, exceeds a certain value
such as β & 0.05. For sufficiently small values of β, however,
particles initially directed toward the host star (which seems
reasonable since the gas or dust emission should commence
on the heated hemisphere) will largely fall toward the host star
until the grains are pushed into orbit by coriolis forces. These
particles will form a “leading tail” in the sense that the motion
will be in front of the planet.
The values of β for the particles are computed according to
β ≃
L∗σ
4πcGM∗µ
≡
3 L∗ h σ(a, λ)i
16πcGM∗ρa
(3)
where L∗ and M∗ are the luminosity and mass of the host star,
a is the particle radius, σ is the particle cross section at wave-
length λ, h σi is the dimensionless cross section in units of
πa2 averaged over the stellar spectrum (see, e.g., Kimura et
al. 2002), and µ and ρ are the mass and material density of
the dust grains. Since, for low-mass main-sequence stars the
luminosity scales roughly as M4
∗, we find that for a fixed ma-
terial in the grains, β scales as
M3
∗ h σ(a, λ)i
(4)
β ∝
a
We evaluate the dimensionless spectrum-averaged cross sec-
tion with a Mie scattering code (Bohren & Huffman 1983) for
different assumed material indices of refraction (see Croll et
al. 2014). Plots of β(a) for several different assumed compo-
sitions are shown in Fig. 14. They are also compared against
representative β(a) curves for KIC 1255b and KOI 2700b.
The latter two systems have β(a) curves that are essentially
a factor of 2 higher than for K2-22b, largely due to the lower
luminosity per unit mass of the latter host star.
Expression (4) raises the interesting prospect that for low
mass stars, the value of β will be sufficiently small that out-
flowing particles passing through their Roche lobes will be
immune to radiation forces and act much in the same way as
FIG. 15.— Simulated dust tails. The dust particles are launched with some-
what more than the escape speed from the planet within a 30◦ cone centered
on the direction of the host star. The particle sizes are chosen via Monte Carlo
means from a power-law distribution, and the value of β for each particle is
calculated from the system parameters (see text). There are 50,000 particles
in each simulation, and each one is assumed to have an exponentially decay-
ing cross section (see App. C of Rappaport et al. 2014) with a time constant of
104 s. Only radiation and gravitational forces are included. The color coding
is proportional to the logarithm of the dust particle density with white-red the
largest to blue-purple the lowest. The dust tails are shown in the rest frame
of the orbiting planet (implicitly moving downward in the frame). The top
panel is a view of the orbit and dust tail from the orbital pole, and clearly
shows both a leading and a trailing tail. The bottom panel is a view from the
orbital plane as the dust emitting planet crosses the disk of the host star. The
middle panel results from arbitrarily multiplying each calculated value of β
by a factor of 4, the net effect of which is to eliminate the leading tail.
K2-22b
17
Roche-lobe overflowing material.
We next carried out a large number of simulations of dust
particles ejected from a planet and initially moving in vari-
ous directions with different speeds. For most of our simu-
lations, we took the direction of the particles to be uniformly
distributed within a cone of 30◦ radius and centered in the
direction of the host star. The velocities were somewhat arbi-
trarily taken to be one half the escape speed from the surface
of the planet’s Roche lobe to infinity in the absence of the host
star, i.e.,pGMp/2RL where Mp is the planet’s mass and RL is
its Roche-lobe radius. This is the minimum that can be con-
templated for a Parker-wind type outflow. For each dust grain,
a radius, a, is chosen from an assumed particle size distribu-
tion via Monte Carlo methods. We considered a power-law
differential size distribution for the dust grains with a slope of
- 2, as illustrative, with the maximum and minimum sizes in
this distribution taken to be 1 µm and 1/20 µm, respectively.
The luminosity of the host star is taken to be that of a main-
sequence M-K star of 0.6 M⊙ (see Table 3). For purposes of
the numerical calculations of the particle dynamics only, we
used a simple analytic fit to the plots shown in Fig. 14 of the
form:
β ∝
(5)
h σ(a, λ)i
a
c1 + c2a3
1 + c3a4
∝
where the c’s are constants to be fit, and which depend on L∗,
M∗, and the dust composition.
The results of our particle dynamics simulations are shown
in Fig. 15. The top panel represents the trajectories of all par-
ticles regardless of their size or corresponding value of β. The
dust density is assumed to decay exponentially in time, with a
time constant of 104 sec. Note that there is a forward moving
“leading tail” in addition to a bifurcated trailing tail. The inner
of the two trailing tails arises from particles with small values
of β that are launched with a substantial velocity component
in the forward direction. In the bottom panel of Fig. 15 we
see the two tails from the perspective of an observer viewing
an equatorial transit; there are far more particles in the lead-
ing tail. For comparison, we show in the middle panel a case
where the computed value of β was arbitrarily multiplied by a
factor of 4; all other parameters remained the same. Note that
the result is a purely trailing dust tail as one would have in a
solar system comet. This latter exercise ensures that all the
particles have a β that exceeds a critical value of ∼0.02, guar-
anteeing that the particles will go into a trailing tail. The bot-
tom line is that dust-emitting planets around luminous stars
should have predominantly trailing tails while the reverse is
true for low-luminosity stars.
In all of these calculations, we have ignored possible ram
pressure forces on the dust grains due to a stellar wind from
the host star. For a justification of why this may be a good
approximation, see Appendix A of Rappaport et al. (2014).
9.3. Model Transits
9.3.1. Idealized Transit Models
The transit profiles expected for a planet with a trailing
comet-like tail are already sufficiently complicated compared
to a conventional hard-body planet transit. They include such
issues as the attenuation caused by the (unknown) absorp-
tion profile, typically characterized by an exponential scale
length32; possible forward scattering which is influenced by
32 For a discussion of why the dust tail may fall off approximately expo-
nentially, including the effects of dust sublimation on a timescale of hours,
FIG. 16.— Schematic transit profiles for a range of dust tail parameters.
θ is the orbital phase of the planet, and θ∗ is the half angle subtended by
the host star as seen from the planet. Top panel: Trailing dust tail only. The
exponential tail length, ℓ, is color coded and is expressed in units of the radius
of the host star. No forward scattering is included. Note how the shorter the
tail length, the smaller is the post-transit depression. Middle panel: Same
as for the top panel, but in this case a contribution from forward scattering
has been included. There is now a noticeable pre-transit bump due to the
forward scattering, but only for the case of short dust tails is there a barely
perceptible post-transit bump; it is typically suppressed by the post-transit
depression. Bottom panel: Here, there is both a trailing and a leading dust
tail. The exponential scale lengths, ℓ, of both dust tails are assumed to be the
same, but the leading tail is taken to have twice the dust density, ρ, at every
corresponding angular distance, as the trailing tail. For long tail lengths there
are no pre- or post-transit bumps due to forward scattering, as they are pulled
below the discernible level by the corresponding depressions caused by direct
attenuation from the tails. By contrast, for the case of short tail lengths, i.e.,
. 0.2 Rstar, there is both a pre- and a post-transit bump, with the latter being
slightly larger.
18
Sanchis-Ojeda et al. 2015
star as seen from the planet. In that case, the effective angular
scattering pattern from a single grain is somewhat larger than
that of the angular size of the host star (i.e., ∼17◦). In the bot-
tom panel of Fig. 16 the transit profiles are for the case where
there is both a trailing and a leading dust tail. The exponen-
tial lengths of both dust tails have been taken to be the same
(i.e., ℓ1 = ℓ2), but the leading tail is taken to have twice the op-
tical thickness everywhere as the trailing tail. For large ℓ there
are no pre- or post-transit bumps, as they are pulled below the
discernible level by the corresponding depressions caused by
the tails. By contrast, for the case of short tail lengths, i.e.,
ℓ . 0.2 R∗, there is both a pre- and a post-transit bump, with
the latter being slightly larger.
9.3.2. Model Fit to the K2-22b Transit
FIG. 17.— Best fitting model transit profile. The black histogram represents
the observed transit (see Fig. 6). The red curve is the best fitting model. The
green and blue curves represent the direct absorption and forward scattering
components, respectively. Note how the forward scattering causes the peaks
on either side of the transit (see also Budaj 2013).
the grain size and composition (which may be changing along
the tail as different materials sublimate at different rates); con-
volution of the scattering phase function with the finite angu-
lar profile of the host star; and another convolution with the
density profile of the tail. When adding the possibility of both
a trailing and leading tail, as we have postulated in this work,
the situation quickly becomes even more complicated. In an
attempt to keep the numbers of free parameters to a minimum,
we have adopted the following seven-parameter model: an ex-
ponential attenuation profile of the tail in both directions, each
with its own scale length, ℓ; a relative density between the two
tails, ρ1/ρ2; the size of the dust grains, a; and an overall nor-
malization factor that yields the correct mean transit depth.
In addition to these parameters associated with the dust tails,
there is also an impact parameter for the transiting planet, and
the time of orbital phase zero. We assume that the hard body
of the planet itself does not make a significant contribution to
the transit profile.
We illustrate in Fig. 16 what some of the possible transit
profiles might look like with just two of these parameters
varying. First, we consider only a trailing dust tail, and we
vary the exponential scale length (ℓ, in units of the host star’s
radius). These transit profiles are shown in the top panel of
Fig. 16. The tails’ exponential scale length, ℓ, is color coded
to help in distinguishing the different curves. No forward
scattering is included in this case. Note how the shorter ℓ is,
the smaller is the post-transit depression. The latter becomes
quite noticeable when ℓ for the tail becomes greater than about
30% of the radius of the star. In the middle panel of Fig. 16
the contribution from forward scattering by relatively large
particles (∼0.5 µm) has been included. In all cases there is
a noticeable pre-transit bump due to the forward scattering,
but only for the case of short dust tails (i.e., small ℓ) is there
a perceptible post-transit bump; it is typically suppressed by
the post-transit depression as the dust tail is in the process of
moving off the stellar disk. The grain sizes assumed here are
sufficiently large so that λ/2πa, the characteristic Mie scat-
tering angle, is comparable with the angular size of the host
see Appendix C of Rappaport et al. (2014).
Finally, we have attempted to fit a simple two-tail dust
model to the observed transit profile of K2-22b. Once the
model and its free parameters were selected, we computed the
scattering pattern for the dust particles (taken to be of a sin-
gle fixed size throughout both tails) via a Mie scattering code
(Bohren & Huffman 1983). Then the scattering pattern was
convolved numerically via a 2D integral with the radiation
profile coming from the host star (including limb-darkening).
In turn, this net effective scattering profile was convolved in
1D with the assumed exponential tails to produce the scatter-
ing pattern as a function of orbital phase. The attenuation of
the beam from the host star was simply computed by placing
a double-sided exponential profile in front of a limb-darkened
stellar disk in different longitudinal locations (i.e., as a func-
tion of orbital phase) and at different vertical locations to get
a best estimate of the impact parameter b (in units of the stel-
lar radius). The sum of the absorption and forward scattering
contributions was then added and convolved with the Kepler
long cadence integration time.
We first explored a wide range of parameter space, espe-
cially focusing on the physically interesting parameters: b,
the impact parameter, ℓ1, ℓ2, ρ1/ρ2, the exponential scale
lengths and ratios of optical thickness of the two tails, respec-
tively, and a, the grain size. We also considered models with
a power-law distribution of grain sizes, but these produced
no better fits than using a single grain size. To simplify the
search, we fixed ℓ1 = ℓ2 under the assumption that the dust
sublimation time and the speed away from the planet is sim-
ilar in the two tails. However, we allowed ρ1/ρ2 to be a free
parameter since the amount of dust in the two tails could be
quite different. From this broad search, we were able to draw
several interesting conclusions. (1) The best fits, by only a
small margin, are found for the single (leading) tail models,
i.e., no trailing tail is needed. (2) For single-leading tail mod-
els, the impact parameter is constrained to be 0.42 < b < 0.78
and 0.19 < ℓ2 < 0.48 R∗ (both 90% confidence limits). (3) For
two-sided tail models the allowed ratio of optical thicknesses
is constrained to be ρ1/ρ2 < 0.5 (assuming ℓ1 = ℓ2). (4) The
best-fitting single grain-size models have 0.3 < a < 0.5 µm.
The final parameter search and error estimation were done
with an MCMC routine. These results are summarized in Ta-
ble 4.
The overall best-fitting model is for b = 0.65 and ℓ2 =
0.32 R∗. The results are shown in Fig. 17. The black his-
togram is the same average transit profile as is shown in Fig. 6.
The red curve is the overall fit to the transit profile, includ-
ing the convolution with the LC integration time. The overall
model transit profile is comprised of two components: the di-
rect attenuation of the flux from the host star (green curve)
K2-22b
19
and the forward scattering from the dust tail back into the
beam directed toward the observer (blue curve). Note how
the forward scattering produces small bumps both before and
after the transits, with the larger bump following the egress.
9.4. Inferred Dust Sublimation Timescale
The dust sublimation timescale depends on the equilibrium
temperature (Teq), the mineral composition, and the size of
the dust grains (see, e.g., Kimura et al. 2002; Appendix C
of Rappaport et al. 2014). Teq for the dust grains in K2-22b
is necessarily somewhat uncertain, ranging between ∼1500
K and 2100 K, depending on exactly how it is computed (to
within unknown factors of order unity). However, if we take
Teq = 1700 K as a typical grain equilibrium temperature, then
the lifetime of a 1 micron grain composed of Fe, SiO, or
fayalite is a matter of a few tens of seconds. On the other
hand, minerals such as enstatite, forsterite, quartz, and corun-
dum, would have sublimation lifetimes of 1/2, 3, 15, and 30
hours, respectively. SiC and graphite could last for a very
much longer time than any of these. (All the dust parame-
ters for the above estimates, including the heat of sublima-
tion, are taken from the conveniently tabulated values of van
Lieshout et al. 2014). Thus, there is no one “expected” dust
sublimation timescale that we can anticipate. If on the other
hand, we believe that the dust tail is only a few degrees of
the orbit in length (see Table 4), and that β ≈ 0.05, then the
“inferred” lifetime is ∼ 1 hour (see Eqn. (4) and (6) of Rap-
paport et al. 2014 for an explanation), and therefore the dust
composition might resemble enstatite or forserite. However,
one should not draw too many conclusions from this since
the actual tail length depends on a combination of numerous
(uncertain) parameters such as the particle sizes (and indeed
the size distribution), the parameter β, the actual equilibrium
temperature, and so forth. Nonetheless, we can say that min-
erals such as enstatite or forsterite do seem consistent with the
observations.
9.5. Dust-on-Dust Collisions
We note that there is an interesting possibility for the dust
streaming from the planet to interact with dust that has already
been orbiting for a substantial while – provided that the subli-
mation lifetime will allow the dust to survive for a sufficiently
long time. The time required for orbiting dust in a trailing
dust tail (i.e., ‘outer-track orbits’ with 0.05 & β & 0.02) to
meet up again with the planet is approximately Porb/(2β) (see
Eqn. (4) of Rappaport et al. 2014); for K2-22b this amounts to
10 - 25 Porb (or ∼4 - 10 days). For particles with β . 0.02 the
orbits likely lie inside the planet’s orbit (‘inner-track orbits’)
with radii between ∼94% and 98% of d. This leads to synodic
orbital periods for those dust particles of between ∼10 and 30
Porb, fairly similar to the outer-track orbits. Thus, any orbit-
ing dust that might collide with newly emitted dust must last
for 4 - 10 days. The dust grains on the inner-track orbits have
speeds in the rest frame of the planet that range from 0 - 6 km
s- 1 (prograde), while the corresponding range of outer-track
orbits is 0 - 25 km s- 1 (retrograde). Thus, the inner track par-
ticles would catch up to, and collide with, newly emitted dust
grains with only a few km s- 1 relative velocity. By contrast,
the outer track particles could collide with newly emitted dust
at relatively high speeds of more than 10 km s- 1. At such
speeds, a collision would deposit a mean energy of ∼1 eV per
atomic mass unit within the dust grain. This seems likely to
completely destroy most dust particles that happen to collide
on the outer track.
Additionally, any orbiting dust grains which are moving
quickly may catch up, and collide, with other more slowly
orbiting dust. Again, the relative speeds of any colliding dust
particles on the inner track would likely be of order a km s- 1
and would not necessarily destroy the dust, whereas the re-
verse is true for the outer-track dust grains.
We can make a crude estimate of the collision frequency
of a dust grain that is orbiting the host star. Take the mean
grain number density along the observer’s line of sight during
a typical transit to be n0, and the collision cross section to be
σ. In that case, the collision mean free path is ℓ = 1/(n0σ).
If we take the radial thickness of the dust tail that is causing
the transits to be ∆r, then ∆r ≃ 1/(n0σ) if both the optical
cross section for extinction and the collision cross section are
approximately equal (i.e., the geometric cross section appro-
priate for larger particles), and if the optical depth of the dust
cloud during transits is of order unity. (The latter is necessary
since the dust tail is thin in the vertical direction and covers
only a small fraction of the disk of the host star (see middle
panel of Fig. 15). Using the above expression, we can esti-
mate the mean collision time, τcoll, for particles orbiting in the
dust disk
τcoll ≈
1
2π
∆r
d
vorb
vrel
n0
hni
Porb
(6)
where vorb and vrel are the mean orbital speed in inertial space
and the mean relative speeds of the particles in a dust disk, and
hni is the mean density in dust grains around the azimuth of
the dust disk. These factors are extremely uncertain, but as an
illustrative example, if we take ∆r/d ≈ 0.05, vorb/vrel ≈ 10,
and n0/hni ≈ 100, then τcoll is of order several times Porb.
What is the fate of dust particles that collide but are not
destroyed in the collision? There could be locations in the
dust disk where the dust is slowed down by the collisions and
the density builds up. In principle, these regions of enhanced
density could be approximately fixed in the rotating frame of
the planet and host star. However, since we see only a single
transit-like feature during an orbital period, we conclude that
such pile-up of dust is not significant anywhere, with the pos-
sible exception of near the planet where the dust density is the
highest.
A careful examination of what the effects are of dust-dust
collisions in this system or, for that matter, KIC 1255b and
KOI 2700b are much beyond the scope of this paper, and we
leave that study to another work.
10. SUMMARY AND CONCLUSIONS
We have reported the discovery in the K2 Field-1 data of
a new planet that is likely the third example of a planet dis-
integrating via the emission of dusty effluents. The planet is
in an ultrashort 9.1457 hour orbit about an M star. The ev-
idence we presented for the presence of a dust tail includes
erratically and highly variable transit depths ranging from
. 0.14% to 1.3% in both the K2 data as well as in follow-up
ground-based observations from five different observatories.
The folded orbital light curve from the K2 data exhibits a clear
post-egress ‘bump’ and a much less significant, but plausible,
pre-ingress ‘bump’, that are not found in conventional hard-
body transits. There is no post-egress depression in the flux
as was seen in KIC 1255b or in KOI 2700b. On at least one
observation with the GTC, the transit depths were distinctly
wavelength dependent with the transits ∼25% shallower at
840 nm than at 630 nm. While this goes in the same direc-
20
Sanchis-Ojeda et al. 2015
tion as limb-darkening effects (see, e.g., Knutson et al. 2007;
Claret & Bloemen 2011), we argue that the magnitude of the
limb-darkening effect explains only a small fraction of what
is observed. Furthermore, the variable behavior of the color-
dependent transit depths strongly suggests an origin in dust
scattering. This requires relatively non-steep power-law par-
ticle size distributions with Γ ≃ 1 to 3 with maximum sizes in
the range of 0.4 - 0.7 µm.
The host star is an M star with Teff ≃ 3800 K. There is a
companion star 1.9′′ away with Teff ≃ 3300 K. We infer a dis-
tance to the system of 225 ± 50 pc, which, in turn, implies a
projected physical separation of the two stars of ∼430 AU.
At this distance, the companion star may have been incidental
to the formation of the planet, but perhaps was instrumental
in driving it toward the host star via Kozai-Lidov cycles with
tidal friction (Kozai 1962, Lidov 1962; Kiseleva et al. 1998;
Fabrycky & Tremaine 2007).
The minimum transit depth of . 0.14% sets an upper limit
to the size of the underlying hard-body planet of 2.5 R⊕ for
an assumed radius of the host star of 0.57 ± 0.06 R⊙. This
is consistent with the low surface gravity that is required to
drive off metal vapors that could condense into dust (Rappa-
port et al. 2012; Perez-Becker & Chiang 2013). In fact, it
seems likely if the dust-emitting scenario we report for K2-
22b is correct then Mars, Mercury, or even lunar sized bodies
with surface gravities of 1/6 to 1/3 that of Earth are to be pre-
ferred.
We find that the dust tail of K2-22b has two properties that
are distinct from those of KIC 1255b33 or KOI 2700b: (1)
a leading dust tail (as opposed to a trailing one); and (2) a
characteristic (e.g., exponential) scale length for the dust tail
that is . 1/2 R∗. The former requires dust transported to about
twice the planet’s radius in the direction of the host star until
it effectively overflows its Roche lobe, thereby going into an
orbit that is faster than that of the planet. It is also necessary
to have β (the ratio of radiation pressure forces to gravity) be
. 0.02 which would be the case for very low luminosity host
stars in combination with very small (. 0.1 µm) or very large
(& 1 µm) dust particles (see Fig. 14). The shorter dust tails
could result from a combination of very low values of β and
a short dust-grain sublimation time of . an hour.
Mass loss rates in the form of high-Z material from this
planet are likely to be ≈ 1.5 × 1011 g s- 1 implying a lifetime
of 20 - 70 Myr, for planet masses in the range of the moon’s
to Mercury’s mass.
The scenario described above for explaining the various
features of the transit light curve and its variability is not en-
tirely self-consistent. The requirement for particles to enter a
leading tail is β . 0.02 µm which implies particle sizes of ei-
ther a . 0.1 µm or a & 1 µm. At the same time, the post-transit
bump requires a forward scattering peak that is comparable to
the angular size of the host star. This corresponds to particle
sizes of ∼ 1/2 µm. Finally, the color dependence of the tran-
sits observed with the GTC implies a non-steep power-law
size distribution with a maximum size of ∼1/2 µm. Thus,
larger particles could account for all three of these observa-
tional pieces of evidence—but only if a substantial fraction
of the particles are large (e.g., ∼ 1 µm). This requirement for
such large particles seems somewhat unusual in comparison
33 We note that Bochinski et al. (2015) were able to measure a small dif-
ference in transit depths in KIC 1255b between g’ and z’ bands, and thereby
concluded that the dusty effluents in this object contained a component of
larger grains in the range of 0.25 - 1 µm.
TABLE 6
COMPARISON OF DUSTY PLANETS
Parameter (units)
EPIC 2016371 KIC 1255b2
KOI 2700b3
21.84
0.031 - 0.053
13
yes
no
slowly
5.9 ± 0.4
9.146
0 - 1.3
highly
3.3 ± 0.2
15.68
0 - 1.4
highly
4.3 ± 0.4
Porb (hr)
Depth (%)
Variability
d/R∗
θ∗ (deg)
Pre-bump
Post-bump
θtail (deg)4
θtail/θ∗
Thost (K)
Teq (K)5
βmax
Mdust (1010 g/s)
1 Results from this work.
2 Values from Rappaport et al. (2012); Brogi et al. (2012); Budaj
(2013); van Werkhoven et al. (2014).
3 Values from Rappaport et al. (2014).
4 Approximate exponential tail length in degrees of orbital phase.
5 Teq ≡ TeffpR∗/d.
17
weak
yes
. 8
. 0.5
3830
2100
0.05
20
10
...
no
∼24
∼2.4
4435
1850
0.07
0.15
∼10 - 15
∼0.77 - 1.1
4300
2100
0.1
20
with other known astrophysical collections of dust such as in
the ISM (e.g., Mathis et al. 1977; Bierman & Harwitt 1980),
Solar-system comet tails (e.g., Kelley et al. 2013), Io (e.g.,
Krüger et al. 2003a; 2003b), and the Earth’s atmosphere (e.g.,
Liou 2002; Holben et al. 1998).
The host star K2-22 shows photometric modulations with
an amplitude of approximately 1% and a timescale of 15 days.
It is interesting to note that the other two host stars to disin-
tegrating planets also display photometric modulations at the
several percent level, likely associated with star spots. It has
been suggested that stellar activity may play an important role
in modulating the process that generates the dust (Kawahara
et al. 2013; Croll et al. 2015). The large spot modulations for
the host stars of the candidate disintegrating planets is sugges-
tive and warrants further investigation.
Finally, we summarize in Table 6 some comparative prop-
erties of the three known ‘disintegrating planets’. Hopefully,
further patterns of similarity will emerge as more of these ob-
jects are discovered.
Discoveries of new disintegrating planets in the upcoming
K2 fields could be potentially quite important. If found, these
candidates are likely to orbit brighter stars than the host stars
of the examples discovered to date. It would also be inter-
esting to discover them orbiting a richer variety of host stars,
especially since we have inferred from K2-22b that the low lu-
minosity of the host star plays an important role in determin-
ing the trajectories of the dust flowing from these very special
planets. Future follow-up observations with ground-based
telescopes, both in photometric and spectroscopic modes, are
likely to provide us with further information about the pro-
cess that generates the dust emission, and give us better in-
sights into the relevant physical processes in these extreme-
environment systems.
We thank Joshua Pepper and Smadar Naoz for helpful com-
ments and an anonymous referee for suggestions that greatly
improved the presentation of the material. We thank Allyson
Bieryla and Dave Latham for helping us with the FLWO ob-
servations of the object. We are grateful to Evan Sinukoff,
Erik Petigura, and Ian Crossfield for observing this target
with Keck/HIRES. We thank Yi Yang for helping with the
HSC images, and Takuya Suenaga for volunteering his time
K2-22b
21
to do so. The help from the Subaru Telescope staff is greatly
appreciated. We acknowledge Tsuguru Ryu for his support
on the Okayama transit observations. T.H. is supported by
Japan Society for Promotion of Science (JSPS) Fellowship
for Research (No. 25-3183).
I.R. acknowledges support
from the Spanish Ministry of Economy and Competitiveness
(MINECO) and the Fondo Europeo de Desarrollo Regional
(FEDER) through grant ESP2013-48391-C4-1-R. We extend
special thanks to those of Hawaiian ancestry on whose sa-
cred mountain of Mauna Kea we are privileged to be guests.
Without their generous hospitality, the Keck observations pre-
sented herein would not have been possible. This work was
performed, in part, under contract with the Jet Propulsion
Laboratory (JPL) funded by NASA through the Sagan Fel-
lowship Program executed by the NASA Exoplanet Science
Institute. This article is partly based on observations made
with the Nordic Optical Telescope operated by the Nordic Op-
tical Telescope Scientific Association and the Gran Telesco-
pio Canarias operated on the island of La Palma by the IAC at
the Spanish Observatorio del Roque de los Muchachos. This
research has been supported by the Spanish MINECO grant
number ESP2013-48391-C4-2-R. The Infrared Telescope Fa-
cility is operated by the University of Hawaii under contract
NNH14CK55B with the National Aeronautics and Space Ad-
ministration. FM acknowledges the support of the French
Agence Nationale de la Recherche (ANR), under the program
ANR-12-BS05-0012 Exo-atmos.
REFERENCES
Aigrain, S., Hodgkin, S.T., Irwin, M.J., Lewis, J.R., & Roberts, S.J. 2015,
MNRAS, 447, 2880
Armstrong, D. J., Kirk, J., Lam, K. W. F., et al. 2015, A&A, 579, A19
Armstrong, D. J., Santerne, A., Veras, D., et al. 2015, arXiv:1503.00692
Barnes, S. A. 2007, ApJ, 669, 1167
Batalha, N. M., Borucki, W. J., Bryson, S. T., et al. 2011, ApJ, 729, 27
Bierman, P. & Harwit, M. 1980, ApJ, 241, L105
Bochanski, J. J., West, A. A., Hawley, S. L., & Covey, K. R. 2007, AJ, 133,
Bochinski, J.J., Haswell, C.A., Marsh, T.R., Dhillon, V.S, & Littlefair, S.P.
531
2015, ApJ, 800, L21
Bohren, C. F., & Huffman, D. R. 1983, New York: Wiley, 1983.
Borucki, W.J., Koch, D., Basri, G., et al. 2010, Sci, 327, 977
Bouchy, F., Queloz, D., Deleuil, M., et al. 2008, A&A, 482, 25
Boyajian, T.S., von Braun, K., van Belle, G. et al. 2012, ApJ, 757, 112
Brogi, M., Keller, C. U., de Juan Ovelar, M., Kenworthy, M. A., de Kok, R.
J., Min, M., & Snellen, I. A. G. 2012, A&A, 545, L5
Budaj, J. 2013, A&A, 557, A72
Chubak, C., Marcy, G., Fischer, D.A., Howard, A.W., Isaacson, H., Johnson,
J.A., & Wright, J.T. 2012, arXiv: 1207.6212
Claret, A. & Bloemen, S. 2011, A&A, 529, A75
Croll, B., Rappaport, S., & DeVore, J. 2014, ApJ, 786, 100
Croll, B., Rappaport, S., & Levine, A. M. 2015, MNRAS, 449, 1408
Crossfield, I. J. M., Petigura, E., Schlieder, J. E., et al. 2015, ApJ, 804, 10
Cushing, M.C., Vacca, W.D., & Rayner, J.T. 2004, PASP, 116, 362
Delfosse, X., Forveille, T., & Ségransan, D. et al. 2000, A&A 364, 217
Delorme, P., Collier Cameron, A., Hebb, L., et al. 2011, MNRAS, 413, 2218
Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935
Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298
Foreman-Mackey, D., Montet, B. T., Hogg, D. W., et al. 2015, ApJ, 806, 215
Frandsen, S., & Lindberg, B. 1999, Astrophysics with the NOT, 71
Fukui, A., Narita, N., Tristram, P.J., et al. 2011, PASJ, 63, 287
Fukui, A., Kawashima, Y., Ikoma, M., et al. 2014, ApJ, 790, 108
Gandolfi, D., Alcalá, J. M., Leccia, S., et al. 2008, ApJ, 687, 1303
Gillon, M., Jehin, E., Magain, P., et al. 2011, in European Physical Journal
Web of Conferences, Vol. 11, European Physical JournalWeb of
Conferences, 6002
AA82
381
Holben, B., Eck, T. F., Slutsker, I., et al. 1998, Rem. Sens. Environ., 66, 1
Holman, M. J., Winn, J. N., Latham, D. W., et al. 2006, ApJ, 652, 1715
Howard, A.W., Johnson, J.A., Marcy, G.W. 2010, ApJ, 721, 1467
Howard, A. W., Sanchis-Ojeda, R., Marcy, G. W., et al. 2013, Nature, 503,
Howell, S.B., Sobeck, C., Haas, M., et al. 2014, PASP, 126, 398
Husser, T.-O., Wende-von Berg, S., Dreizler, S., Homeier, D., Reiners, A.,
Barman, T., & Hauschildt, P.H. 2013, A&A, 553, 6
Jackson, B., Barnes, R., & Greenberg, R. 2009, ApJ, 698, 1357
Jehin, E., Gillon, M., Queloz, D., et al. 2011, The Messenger, 145, 2
Jenkins, J.M., Caldwell, D.A., Chandrasekaran, H., et al. 2010, ApJ Lett,
Kawahara, H., Hirano, T., Kurosaki, K., Ito, Y., & Ikoma, M. 2013, ApJ,
713, L 87
776, LL6
529
Kovács, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369
Krüger, H., Horányi, M. & Gün, E., 2003a, GeoRL, 30.1058
Krüger, H., Geissler, P. & Horányi, M. et al. 2003b, GeoRL, 30.2101
Knutson, H.A., Charbonneau, D., Noyes, R.W., Brown, T.M., & Gilliland,
R.L. 2007, ApJ, 655, 564.
Kozai, Y., 1962, AJ, 67, 591
LaCourse, D. M., Jek, K. J., Jacobs, T. L., et al. 2015, MNRAS, 452, 3561
Lantz, B., Aldering, G., Antilogus, P., et al., 2004, SPIE 5249, 146
Lázaro, C., Arévalo, M. J., & Almenara, J. M. 2015, ??jnlNew A, 34, 139
Lépine, S., Hilton, E. J., Mann, A. W., et al. 2013, AJ, 145, 102
Lissauer, J.J., Marcy, G.W., Bryson, S.T., et al. 2014, ApJ Lett, 784, L44
Lidov, M. L., 1962, PlanSS, 9, 719
Liou. K.N. 2002, An Introduction to Atmospheric Radiation, Second
Edition, Academic Press, San Diego, CA
Lund, M. N., Handberg, R., Davies, G. R., Chaplin, W. J., & Jones, C. D.
2015, ApJ, 806, 30
Maldonado, J., Affer, L., Micela, G., et al. 2015, A&A, 577, A132
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Mann, A.W., Gaidos, E., & Aldering, G. 2011, PASP 123, 1273
Mann, A.W., Gaidos, E., & Ansdell, M. 2013, ApJ, 779, 188
Mann, A. W., Feiden, G. A., Gaidos, E., Boyajian, T., & von Braun, K.
2015, ApJ, 804, 64
Mathis, J. S., W. Rumpl, & Nordsieck, K. H. 1977, ApJ, 217, 425
McQuillan, A., Aigrain, S., & Mazeh, T. 2013, MNRAS, 432, 1203
Miyazaki, S., Komiyama, Y., Nakaya, H., et al. 2012, Proc. SPIE, 8446,
84460Z
Montet, B. T., Morton, T. D., Foreman-Mackey, D., et al. 2015, ApJ, 809, 25
Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015, ApJS, 217, 31
Murgas, F., Pallé, E., Zapatero Osorio, M. R., et al. 2014, A&A, 563, AA41
Ofir, A. 2014, A&A, 561, AA138
Pepe, F., Cameron, A. C., Latham, D. W., et al. 2013, Nature, 503, 377
Perez-Becker, D., & Chiang, E. 2013, MNRAS, 433, 2294
Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231
Rappaport, S., Levine, A., Chiang, E., et al. 2012, ApJ, 752, 1
Rappaport, S., Sanchis-Ojeda, R., Rogers, L. A., Levine, A., & Winn, J. N.
2013, ApJ, 773, L15
Rappaport, S., Barclay, T., DeVore, J., Rowe, J., Sanchis-Ojeda, R., & Still,
Rayner, J.T., Toomey, D.W., Onaka, P.M., Denault, A.J., Stahlberger, W.E.,
Vacca, W.D., Cushing, M.C., & Wang, S. 2003, PASP, 115, 362
Rayner, J. T., Cushing, M. C., & Vacca, W. D. 2009, ApJS, 185, 289
Reiners, A., & Mohanty, S. 2012, ApJ, 746, 43
Rowe, J.F., Bryson, S.T., Marcy, G.W., et al. 2014, ApJ, 784, 45
Sanchis-Ojeda, R., Rappaport, S., Winn, J. N., et al. 2013, ApJ, 774, 54
Sanchis-Ojeda, R., Rappaport, S., Winn, J. N., et al. 2014, ApJ, 787, 47
Schechter, P.L., Mateo, M., & Saha, A. 1993, PASP, 105, 1342
Schlaufman, K. C., Lin, D. N. C., & Ida, S. 2010, ApJ, 724, L53
Stassun, K.G., Pepper, J.A., Paegert, M., De Lee, N., & Sanchis-Ojeda, R.
Telting, J. H., Avila, G., Buchhave, L., et al. 2014, Astronomische
2014, arXiv: 1410.6379
Nachrichten, 335, 41
Vacca, W. D., Cushing, M. C., & Rayner, J. T. 2003, PASP, 115, 389
Valdes, F., Gupta, R., Rose, J. A., Singh, H. P., & Bell, D. J. 2004, ApJS,
van Werkhoven, T. I. M., Brogi, M., Snellen, I. A. G., & Keller, C. U. 2014,
Gillon, M., Anderson, D. R., Collier-Cameron, A., et al. 2013, A&A, 552,
M., 2014, ApJ, 784, 40
Kelley, M.S., Lindler, D.J., Bodewits, D., et al. 2013, Icarus, 222, 634.
Kimura, H., Mann, I., Biesecker, D.A., & Jessberger, E.K. 2002, Icarus, 159,
152, 251
A&A, 561, 3
Kiseleva, L. G., Eggleton, P. P., & Mikkola, S, 1998, MNRAS, 300, 292
Vanderburg, A., & Johnson, J. A. 2014, PASP, 126, 948
22
Sanchis-Ojeda et al. 2015
Vanderburg, A., Montet, B. T., Johnson, J. A., et al. 2015, ApJ, 800, 59
van Lieshout, R., Min, M., & Dominik, C. 2014, A&A, 572, 76
Winn, J. N., Holman, M. J., Torres, G., et al. 2008, ApJ, 683, 1076
Yanagisawa, K., Shimizu, Y., Okita, K., et al. 2006, Proc. SPIE, 6269,
62693Q
|
1509.02522 | 1 | 1509 | 2015-09-08T20:03:19 | NEOWISE Reactivation Mission Year One: Preliminary Asteroid Diameters and Albedos | [
"astro-ph.EP"
] | We present preliminary diameters and albedos for 7,959 asteroids detected in the first year of the NEOWISE Reactivation mission. 201 are near-Earth asteroids (NEAs). 7,758 are Main Belt or Mars-crossing asteroids. 17% of these objects have not been previously characterized using WISE or NEOWISE thermal measurements. Diameters are determined to an accuracy of ~20% or better. If good-quality H magnitudes are available, albedos can be determined to within ~40% or better. | astro-ph.EP | astro-ph | NEOWISE Reactivation Mission Year One: Preliminary Asteroid
Diameters and Albedos
C. R. Nugent1, A. Mainzer2, J. Masiero2, J. Bauer2, R. M. Cutri1, T. Grav3, E. Kramer2,
S. Sonnett2, R. Stevenson, and E. L. Wright4
Received
;
accepted
1Infrared Processing and Analysis Center, California Institute of Technology, Pasadena,
CA 91125, USA
2Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109 USA
3Planetary Science Institute, Tucson, AZ
4Department of Physics and Astronomy, University of California, Los Angeles, CA 90095,
USA
5
1
0
2
p
e
S
8
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
2
2
5
2
0
.
9
0
5
1
:
v
i
X
r
a
ABSTRACT
We present preliminary diameters and albedos for 7,959 asteroids detected
in the first year of the NEOWISE Reactivation mission. 201 are near-Earth
asteroids (NEAs). 7,758 are Main Belt or Mars-crossing asteroids. 17% of these
objects have not been previously characterized using WISE or NEOWISE thermal
measurements. Diameters are determined to an accuracy of ∼ 20% or better. If
good-quality H magnitudes are available, albedos can be determined to within
∼ 40% or better.
1.
Introduction
Sizes and albedos of asteroids are basic quantities that can be used to answer a
range of scientific questions. A significant number of diameter measurements produce a
size-frequency distribution, which can constrain models of asteroid formation and evolution
(Zellner 1979; Gradie & Tedesco 1982; Tedesco et al. 2002; Bus & Binzel 2002; Masiero
et al. 2011). Asteroid albedos aid the identification of collisional family members (Masiero
et al. 2013; Walsh et al. 2013; Carruba et al. 2013; Milani et al. 2014; Masiero et al. 2015),
and allow for basic characterization of asteroid composition (Mainzer et al. 2011c; Grav
et al. 2012b; Masiero et al. 2014).
Most observations of asteroids are made in visible wavelengths, where flux is dependent
on both size and albedo. Observations in other wavelengths, such as the infrared (e.g.
Hansen 1976; Cruikshank 1977; Lebofsky et al. 1978; Morrison & Lebofsky 1979; Delb´o
et al. 2003, 2011; Matter et al. 2011; Muller et al. 2012, 2013; Wolters et al. 2005, 2008) or
radio (e.g. Ostro et al. 2002; Benner et al. 2015), are needed to determine these quantities
precisely. At present, well-determined diameters and albedos have been measured for less
than a quarter of known asteroids.
2
The infrared NEOWISE project (Mainzer et al. 2011a) has measured diameters and
albedos for ∼ 20% of the known asteroid population, the majority of these measurements to
date (Mainzer et al. 2011b, 2012, 2015; Masiero et al. 2011, 2012; Grav et al. 2011, 2012a;
Bauer et al. 2013). Here, we expand the number of asteroids characterized by NEOWISE,
deriving diameters and albedos for asteroids detected by NEOWISE between December 13,
2013, to December 13, 2014 during the first year of the Reactivation mission.
The NEOWISE mission uses the Wide-Field Infrared Survey Explorer (WISE)
spacecraft, which images the entire sky using freeze-frame scanning from a sun-synchronous
polar orbit (Wright et al. 2010; Cutri et al. 2012). WISE is equipped with a 50 cm telescope
and four 1024x1024 pixel focal plane array detectors that simultaneously image the same
47x47 arc minute field-of-view in 3.4, 4.6, 12 and 22 µm bands, all originally cooled by solid
hydrogen cryogen. WISE scans the sky between the ecliptic poles continuously during its 95
minute orbit. A tertiary scan mirror freezes the sky on the focal planes for 11 seconds while
the detectors are read out, producing a sequence of adjacent images with 7.7 sec exposure
times in 3.4 and 4.6 µm bands and 8.8 sec in 12 and 22 µm bands. The orbit precesses at
an average rate of approximately one degree per day, so that the full sky is covered in six
months.
WISE was launched on December 14, 2009, and began surveying on January 7,
2010. WISE scanned the sky 1.5 times during the 9.5 months while it was cooled by its
hydrogen cryogen. After the hydrogen was depleted, the survey continued as NEOWISE
until February 1, 2011 using the 3.4 and 4.6 µm detectors that operated at near full
sensitivity with purely passive cooling. During the additional four months of “post-cryo”
operations, coverage of the entire inner Main Asteroid Belt was completed, along with
a second complete coverage of the sky. WISE/NEOWISE was placed into hibernation
in mid-February 2011. In this mode, the solar panels were held facing the Sun and the
telescope pointed towards the north ecliptic pole. The telescope viewed the Earth during
3
half of each orbit, resulting in some heating.
The WISE spacecraft was brought out of hibernation in September 2013 and renamed
NEOWISE to continue its mission to discover, track and characterize asteroids through
∼ 2017 (Mainzer et al. 2014). The telescope was restored to near zenith pointing, which
enabled the optics and focal planes to cool passively back to ∼ 74K. Survey operations
resumed on December 13, 2013, with the 3.4 and 4.6 µm detectors operating at a sensitivity
comparable to that during the original WISE cryogen survey (Cutri et al. 2015). The
NEOWISE moniker, the acronym of Near-Earth Object WISE, encompasses both the
archiving of individual images, to allow for the detection of transient objects, and the
extensions of the mission beyond WISE’s original 9-month lifetime.
NEOWISE uses the same survey and observing strategy as the original WISE mission
(Wright et al. 2010). The majority of each orbit is devoted to observations, with only brief
breaks for data transmission and momentum unloading. The spacecraft carries a body-fixed
antenna, and therefore must reorient itself to communicate with the Tracking and Data
Relay Satellite System (TDRSS), which relays the data to Earth. Data transmission
is timed to only interrupt survey coverage near the ecliptic poles, which are observed
frequently. Momentum unloading, which can result in streaked images, is also completed at
this time.
Data processing for NEOWISE uses the WISE Science Data System (Cutri et al. 2015)
that performs instrumental, photometric and astrometric calibration for each individual
set of 3.4 and 4.6 µm exposures obtained by the spacecraft, and detects and characterizes
sources on each exposure. The calibrated images and the database of positions and fluxes
of sources extracted from those images for the first year of NEOWISE survey observations
were released in March 2015 (Cutri et al. 2015).
The WISE Moving Object Pipeline System (WMOPS; Cutri et al. 2012) identifies
4
sources that display motion between the different observations of the same region on the
sky. WMOPS uses the extracted source lists from sets of images to first identify and
filter out sources that appear stationary between individual exposures, and then links
non-stationary detections into sets that exhibit physically plausible motion on the sky.
Generally, objects within 70 AU of the sun move quickly enough to be detected by WMOPS
(Mainzer et al. 2011a, see also Bauer et al. 2013). Those candidate moving objects that
are not associated with known asteroids, comets, planets or planetary satellites are verified
individually by NEOWISE scientists. A minimum of five independent detections are
required for a tracklet (a set of position/time pairs) to be considered reliable. Tracklets for
each verified new candidate object and previously known solar system objects are reported
to the IAU Minor Planet Center (MPC) three times per week. The MPC performs initial
orbit determination, associates the NEOWISE tracklets with known objects, and archives
the NEOWISE astrometry and times in its observation database.
Candidates confirmed by the MPC to be possible new near-Earth objects (NEOs) are
posted to their NEO Confirmation page for prompt follow-up observations by ground-based
observers. Rapid follow-up is essential for NEOWISE NEO candidates because the
NEOWISE arcs are usually short, and the asteroid’s projected positional uncertainties grow
quickly, making reliable recovery difficult after 2-3 weeks. To ensure prompt follow-up,
NEOWISE observations are reported to the MPC less than three days after observations on
board the spacecraft. A NEOWISE candidate discovery has a minimum of 5 observations
over ∼ 3 hours, although typical objects have ∼ 12 observations spanning ∼ 1.5 days.
Targets observed by NEOWISE can pose unique challenges to ground based follow-up
observers. NEOWISE’s orbit allows observations to be made at all declinations, and
observing is independent of lunar phase. Ground-based observers are limited to a fixed
declination range, and must sometimes deal with light from the moon and terrestrial
weather, which can preclude observations. Moreover, NEOWISE discoveries are frequently
5
extremely dark (see Figure 5), often requiring 2-4 m class telescopes to detect them at low
solar elongations.
Observers around the globe (including both amateurs and professionals) have
contributed essential follow-up observations, which are defined here as an observation
of an object within 15 days of its first observation on board the spacecraft. Significant
contributors of follow-up observations are given in Figure 1. The Spacewatch Project
(McMillan 2007) contributes a large share of recoveries in the northern hemisphere. The
Las Cumbres Observatory Global Telescope (LCOGT) Network of robotically operated
queue-scheduled telescopes (Brown et al. 2013) has been an extremely useful resource for
securing detections when weather is poor at a particular site. Additionally, the group led by
D. Tholen using the University of Hawaii 2.2 m and Canada-France-Hawaii 4 m telescopes
has successfully detected the targets with the faintest optical magnitudes in the northern
hemisphere (e.g. Tholen et al. 2014). The NEOWISE team was awarded time with the
DECam instrument on the Cerro Tololo Inter-American Observatory (CTIO) 4m telescope,
which has proven invaluable for the recovery of low albedo objects at extreme declinations
in the southern hemisphere.
We present diameters and albedos for 201 near-Earth asteroids (NEAs) and 7,758
Main Belt and Mars-crossing asteroids detected in the first year of reactivation, between
December 13, 2013, and December 13, 2014. This includes the 38 NEAs discovered by
NEOWISE during those dates.
6
Fig. 1.— Number of follow-up observations by observatories that contributed > 5 observa-
tions during the Year 1 Reactivation. Spacewatch, LCOGT, and Catalina employ multiple
telescopes; their observatory codes have been grouped together. Observatory code 568,
Mauna Kea, is frequently used by the Tholen group.
2. Methods
2.1. Observations
The MPC is responsible for verifying and archiving asteroid astrometry. To obtain the
verified record of objects found by the WMOPS pipeline in the NEOWISE data, we queried
7
the MPC observations files ‘NumObs.txt’ and ‘UnnObs.txt’ for NEOWISE (observatory
code C51) observations between December 13, 2013 and December 13, 2014. This returned
the list of object identifications, along with the observation times and NEOWISE measured
astrometry. This included known objects and WMOPS asteroid discoveries made during
that time.
The NASA/IPAC Infrared Science Archive (IRSA, at http://irsa.ipac.caltech.edu)
NEOWISE-R Single Exposure Source Table was then queried for the fluxes of sources
detected in the NEOWISE data. The list of detections extracted from the MPC files was con-
verted into GATOR format (see http://irsa.ipac.caltech.edu/applications/Gator/GatorAid/irsa/QuickGuidetoGator.htm),
and uploaded into the IRSA interface using a cone search radius of 2 arcsec and a restriction
that times match the MPC-archived observation time to within 2 seconds. This two-step
process of querying both the MPC archive and the NEOWISE-R Single Exposure
(L1b) Source Table ensures that only detections verified both by the NEOWISE object
identification routines and the MPC are used for thermal modeling. While there may be
additional objects in the database that were detected fewer than 5 times, or are just below
the single-frame detection threshold, this method of extracting moving object detections
ensures high reliability, since WMOPS actively works to exclude fixed sources such as stars
and galaxies from tracklets. Sources with fewer than 5 detections or those that fall just
below the single-frame detection threshold will be extracted in future processing.
NEOWISE detections were further filtered using several measurement and image
quality flags. We required detections to have “ph qual” values of “A”, “B”, or “C”, “cc flag”
values of “0”, and “qual frame” values of “10”. The ‘ph qual” flag represents photometric
quality, accepting a value of “C” or higher ensures that the sources was detected with a
flux signal-to-noise ratio > 2. The “cc flag”, or contamination and confusion flag, indicates
whether the source measurement may be compromised due to a nearby image artifact.
By filtering for “cc flag”= 0, we select for sources unaffected by known artifacts. Finally,
8
“qual frame” is an overall quality grade for the entire image in which the source was
detected. We accepted only the best-quality images, those with a score of “10”.
The filtered data from the NEOWISE Single Exposure Source Table are high-quality
source measurements that were found at the times and locations of NEOWISE WMOPS
detections submitted to the MPC. To further guard against the possibility of confusing
a minor planet with fixed background sources such as stars and galaxies, we uploaded
the filtered data to the IRSA catalog query engine, referencing the WISE All-Sky Source
Catalog to determine if any single-frame detections fell within 6.5 arcsec of an Atlas source.
The WISE Source Catalog is generated using multiple independent single exposure images.
Fast-moving solar system objects are suppressed during the construction of the catalog. A
search radius of 6.5 arcsec was chosen as it is the approximate size of the WISE beam in
the 3.4 and 4.6 µm bands.
We required at least three observations with magnitude errors σmag ≤ 0.25 in one band.
The largest main-belt asteroids (MBAs) can saturate the NEOWISE detectors, resulting
in reduced photometric accuracy. Following the prescription laid out in Cutri et al. (2012)
(Section IV.4), we did not consider objects that were brighter than 8.0 mag at 3.4µm and
7.0 mag at 4.6µm. The NEA measurements used in this work are given in Table 1.
9
Table 1. NEOWISE magnitudes for the NEAs modeled in this paper. Given are the time
of the observation in modified Julian date (MJD), and the magnitude in the 3.4µm (W1)
and 4.6µm bands (W2). Non-detections at a particular wavelength represent 95%
confidence limits (Cutri et al. 2012). The aperture radius in arcsec used for aperture
photometry is given under “Aperture”; “0” indicates that the pipeline profile fit
photometry was used. Observations for the first three objects only are shown; the
remainder are available in electronic format through the journal website.
Name
MJD
W1 (mag)
W2 (mag)
Aperture
01566
56795.5373147
01566
56795.668982
01566
56795.8005219
01566
56795.8663555
01566
56795.9321892
01566
56796.1295626
01566
56796.2612299
>16.339
15.340 ± 0.132
15.270 ± 0.133
15.268 ± 0.137
15.590 ± 0.200
14.904 ± 0.097
15.829 ± 0.192
01580
56955.905682
>16.484
01580
56956.037222
01580
56956.431715
01580
56956.5631277
01580
56956.6289614
01580
56956.6946677
01580
56956.7603741
01580
56956.8262078
01580
56956.8919142
01580
56956.9576205
01580
56957.0891606
01580
56957.4179471
01620
56993.9087248
01620
56994.3030911
01620
56994.434504
01620
56994.5659171
01620
56994.5660444
01620
56994.6317509
01620
56994.7631639
01620
56994.8945768
01620
56994.8947042
01620
56994.9604107
01620
56995.0918237
01620
56995.2890705
01620
56995.4204835
>16.124
17.100 ± 0.538
16.951 ± 0.474
>16.168
16.178 ± 0.252
16.442 ± 0.316
>17.166
16.944 ± 0.523
16.206 ± 0.291
16.795 ± 0.397
>17.009
15.427 ± 0.137
15.463 ± 0.154
15.420 ± 0.144
15.596 ± 0.171
16.012 ± 0.212
15.754 ± 0.198
15.513 ± 0.145
15.794 ± 0.216
15.843 ± 0.488
15.241 ± 0.129
15.203 ± 0.109
15.354 ± 0.124
15.411 ± 0.161
10
13.317 ± 0.086
13.287 ± 0.104
13.255 ± 0.157
13.226 ± 0.125
13.395 ± 0.166
13.348 ± 0.102
13.467 ± 0.196
14.033 ± 0.171
14.230 ± 0.156
13.972 ± 0.136
14.158 ± 0.198
14.159 ± 0.157
14.312 ± 0.187
13.976 ± 0.154
13.988 ± 0.209
14.050 ± 0.134
14.282 ± 0.186
14.271 ± 0.180
13.987 ± 0.145
14.075 ± 0.156
14.049 ± 0.200
13.556 ± 0.102
14.375 ± 0.205
14.305 ± 0.221
13.846 ± 0.169
13.807 ± 0.132
14.203 ± 0.228
14.106 ± 0.155
13.988 ± 0.155
13.637 ± 0.140
13.483 ± 0.115
13.768 ± 0.175
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
0
2.2. NEATM
We used the Near-Earth Thermal Model (NEATM) of Harris (1998), following the
implementation of Mainzer et al. (2011b, 2012) for NEAs and Masiero et al. (2011, 2012) for
MBAs and Mars crossers. These results supersede those published in Mainzer et al. (2014).
NEATM is a simple but effective method for determining effective spherical diameters and
albedos (when corresponding visible light observations are available). This model makes
several assumptions, including a spherical, non-rotating body, with a simple temperature
distribution:
T (θ) = Tmax cos1/4(θ)
for 0 ≤ θ ≤ π/2
where θ is the angular distance from the sub-solar point. Tmax is the subsolar
(cid:18)(1 − A)S
(cid:19)1/4
temperature, defined as:
Tmax =
ησ
(1)
(2)
where A is the bolometric Bond albedo, S is the solar flux at the asteroid, η is termed
the beaming parameter, is the emissivity, and σ is the Stefan-Boltzmann constant. The
beaming parameter η accounts for any deviation between the actual asteroid and the model.
Changes in η can account for a host of factors including non-spherical shapes, the presence
of satellites, variations in surface roughness or thermal inertia, uncertainties in emissivity,
high rates of spin, changes in surface temperature distributions due to spin pole location,
or the imprecise assumption that the object’s night-side has zero thermal emission (a factor
Table 1—Continued
Name
MJD
W1 (mag)
W2 (mag)
Aperture
01620
56995.8147225
15.912 ± 0.223
14.289 ± 0.205
0
11
that is most relevant for objects observed at high phase angles). Some of these factors
accounted for in the beaming parameter are degenerate. For example, a slow-rotating
object will have a heat distribution similar to a faster rotating object that has a lower
thermal inertia. For this simple model, beaming accounts for the changes in temperature
distribution due to these effects that cannot be otherwise separated.
Observations were divided into apparitions of 10 days, and the NEATM model
was fitted to each individual apparition. These shorter apparitions allowed for fits to
widely-spaced apparitions or, for NEAs, over changing phase angles. Given that the
NEOWISE observational cadence generally results in an object being detected over ∼ 1.5
days, sometimes with an additional epoch of observations ∼ 3 − 6 months later, we chose to
divide observations separated by > 10 days for separate fitting to account for large changes
in object distances and viewing geometries.
NEATM spheres were approximated by a faceted polygon with 800 facets. Individual
facet temperature was determined following Equation 1, and then color corrected following
Wright et al. (2010). Observed thermal flux for each facet was computed, as was flux from
reflected sunlight. The integrated flux from the object was determined, accounting for
viewing geometry, to produce a model magnitude. A least-squares fitting routine compared
modeled to observed magnitudes, and iterated on diameter, albedo, and beaming parameter
until a best fit was found.
Geometric optical albedo pV was computed using absolute magnitude H and slope
parameter G, using values supplied in MPCORB.dat by the MPC. Inaccurate H and G
values will result in inaccurate pV fits. Work by Williams (2012) and Pravec et al. (2012)
found systematic H offsets that vary as a function of H magnitude in data reported to the
MPC. As albedo measurements depend on H and G values, errors in measurement of those
values will propagate to derived albedos.
12
NEATM requires at least one of the NEOWISE wavelengths to be dominated by
thermally emitted light. Some outer main-belt objects observed by NEOWISE were too
cold to have thermally dominated emission at 3.4 or 4.6 µm, and therefore diameters and
albedos for those objects are not reported in this paper. The proportion of reflected vs.
thermally emitted light for NEAs and inner MBAs can be seen in the spectral energy
distribution plots shown in Figure 2. The proportion of thermally emitted flux depends on
albedo, which means that for colder, outer MBAs it is unclear if a wavelength is thermally
dominated until after the fit was performed. Comparison of those results to NEOWISE fits
using 12 micron images and radar data confirmed that the thermal fits were poor, so we did
not include results that had more than 25% reflected light in the 4.6µm band.
We assumed that η was equal to the average value for the object’s population, as
determined by Mainzer et al. (2011b) or Masiero et al. (2011), respectively. For NEAs,
this meant η = 1.4 ± 0.5; for all other asteroids in this paper, η = 0.95 ± 0.25. As shown
in Masiero et al. (2011), although the average η for the main belt is 1.0, the peak of the
histogram is located closer to 0.95, so this value was adopted in this work.
Following the average values determined by Mainzer et al. (2011b) and Masiero et al.
(2011), the ratio of infrared to visible albedo pIR/pV was initially set to 1.6 ± 1.0 for NEAs
and 1.5 ± 0.5 for Mars-crossers and MBAs. Additionally, it was assumed that the albedos
of each band were equal, or p3.4µm = p4.6µm. Although this may be a poor assumption for
objects with red slopes (Grav et al. 2012c), it is necessary to prevent over fitting of the
data.
2.3. Uncertainties
Uncertainties on d, pV , and η (when η was a free parameter) were determined using a
Monte Carlo method. Measured NEOWISE magnitudes, H and G were randomly adjusted
13
Fig. 2.— Comparison of spectral energy distribution for a simulated NEO and inner main-
belt asteroid, each with albedos ranging from pV = 0.06 to pV = 0.5. Thick lines show the
flux from the asteroid as a function of wavelength, which is composed of both thermally
emitted (dashed) and reflected sunlight (dotted) components. NEOWISE bands centered at
3.4 and 4.6 µm are shown as shaded cyan and purple bars, respectively. For NEAs (left),
the 3.4 and 4.6 µm bands are both thermally dominated. For objects in the inner Main Belt
(right), the 3.4 µm band is dominated by reflected light, and the 4.6 µm band is dominated
by thermally emitted light, though the ratio between these components varies with albedo.
within their errors, and the resultant model values of d, pV , and (in appropriate cases) η,
were compared to the best-fit values. This process was repeated 25 times for each object,
and the resultant errors are the weighted standard deviation of the Monte Carlo trials. The
errors quoted in the tables below only include the random component measured through
this MC method, not the systematic offset.
Systematic errors were computed by comparing the match between diameters derived
in this work to radar-derived diameters for the same objects. Albedos were derived from
the radar diameters using the equation
d =
10−H/5
1329√
pV
14
(3)
where d is the diameter in kilometers (for more information, see Harris & Lagerros 2002).
2.4. Objects without visible-wavelength detections
Some MBA and Mars-crossing asteroids had no visible-wavelength measurements
available from the MPC. Unlike NEAs, objects determined to have these orbits by the MPC
are not added to the MPC’s NEO Confirmation Page. Therefore, optical follow-up of these
objects is rare, and usually serendipitous. For objects without reported optical detections,
the H values in MPCORB.dat represent estimates, not measurements, and pV could not
be derived. Since thermally emitted light weakly depends on albedo, d measurements are
reported for these objects. However, lacking targeted follow-up, these objects have short
arcs and relatively large position uncertainties, which can add additional systematic errors
to the derived diameters.
2.5. NEAs
NEAs were examined with particular care. Objects with bad matches between observed
and modeled H values were refitted with a parameter that tightened the constraints
between modeled and observed H. Finally, in some cases an assumption of fixed η = 1.4
produced a poor result. For NEAs with poor fits, beaming was varied between 1.0 and 2.0,
to see if a statistically significant improvement in fit to the observed NEOWISE magnitudes
could be achieved.
3. Results
Results are divided into four tables. As diameters were calculated using different
parameters for the NEAs vs the MBAs and Mars-crossing asteroids, results for these two
15
groups are presented separately. Results are further subdivided between objects that were
characterized previously by the NEOWISE team, and objects that were not. This is because
previously published values likely used the fully cryogenic 12 and 22 µm wavelengths, and
therefore can derive diameters more accurately, to within 10%. Researchers looking for the
best-constrained diameter and albedo measurements should consult previously published
work (Mainzer et al. 2011b; Masiero et al. 2011; Mainzer et al. 2012; Masiero et al. 2012).
However, for those researchers who are interested in diameters and albedos derived from
additional epochs of data provided by the Year One Reactivation results, we also include
the diameters derived for objects using these new data.
Tables 2 and 3 contain the fit diameters and albedos for 173 new and 28 previously
characterized NEAs, respectively. Tables 4 and 5 contain the fit diameters and albedos
for 1,176 new and 6,579 previously characterized MBAs and Mars crossing asteroids,
respectively. Several objects were observed at multiple apparitions; in these cases, results
are presented for each apparition.
16
Table 2. Measured diameters (d) and albedos (pV ) of near-Earth objects not previously
characterized using NEOWISE data. Magnitude H, slope parameter G, and beaming η
used are given. The numbers of observations used in the 3.4 µm (nW 1) and 4.6 µm (nW 2)
wavelengths are also reported, along with the amplitude of the 4.6 µm light curve (W2
amp.).
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
1566
1580
1620
1862
1862
1917
1943
1943
1943
2062
3288
4954
5381
5381
6053
7025
7889
8567
13651
35107
35107
39572
39796
53430
54686
55532
68063
68267
68348
68548
68548
85182
01566
16.90
0.15
01580
14.50
0.15
01620
15.60
0.15
01862
16.25
0.09
01862
16.25
0.09
01917
13.90
0.15
01943
15.75
0.15
01943
15.75
0.15
01943
15.75
0.15
02062
16.80
0.15
03288
15.20
0.15
04954
12.60
0.15
05381
16.50
0.15
05381
16.50
0.15
06053
14.90
0.15
07025
18.30
0.15
07889
15.20
0.15
08567
15.30
0.15
13651
17.60
0.15
35107
16.80
0.15
35107
16.80
0.15
39572
16.50
0.15
39796
15.70
0.15
53430
16.60
0.15
54686
16.50
0.15
55532
16.10
0.15
68063
15.50
0.15
68267
16.90
0.15
68348
14.20
0.15
68548
16.50
0.15
68548
16.50
0.15
85182
17.10
0.15
1.03 ± 0.04
8.55 ± 5.23
1.87 ± 0.05
1.40 ± 0.04
1.26 ± 0.04
4.99 ± 0.14
2.34 ± 0.05
2.30 ± 0.04
2.28 ± 0.05
0.80 ± 0.03
2.49 ± 0.07
9.56 ± 0.24
0.91 ± 0.05
0.94 ± 0.04
3.72 ± 0.08
0.50 ± 0.17
1.68 ± 0.07
2.93 ± 0.07
0.56 ± 0.02
0.91 ± 0.03
1.10 ± 0.28
1.55 ± 0.66
2.13 ± 0.59
1.23 ± 0.32
1.35 ± 0.46
1.31 ± 0.04
2.30 ± 0.07
0.88 ± 0.04
3.51 ± 0.13
1.18 ± 0.04
1.24 ± 0.04
1.03 ± 0.37
0.29 ± 0.05
0.04 ± 0.08
0.29 ± 0.04
0.29 ± 0.04
0.35 ± 0.05
0.20 ± 0.03
0.16 ± 0.02
0.17 ± 0.02
0.17 ± 0.03
0.52 ± 0.10
0.24 ± 0.04
0.18 ± 0.03
0.54 ± 0.07
0.51 ± 0.06
0.14 ± 0.02
0.34 ± 0.23
0.52 ± 0.06
0.16 ± 0.03
0.51 ± 0.11
0.41 ± 0.05
0.28 ± 0.16
0.18 ± 0.16
0.20 ± 0.20
0.27 ± 0.15
0.24 ± 0.19
0.38 ± 0.06
0.21 ± 0.04
0.40 ± 0.05
0.30 ± 0.05
0.32 ± 0.04
0.29 ± 0.03
0.24 ± 0.19
17
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.37
1.40 ± 0.47
1.40 ± 0.39
1.40 ± 0.37
1.40 ± 0.47
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.49
0.24
0.34
0.89
0.41
0.84
0.58
0.22
0.29
0.54
0.82
1.40
1.06
0.49
0.17
0.21
0.58
0.45
0.42
1.22
0.25
0.45
0.42
0.69
1.14
1.02
0.22
0.38
0.29
0.46
0.24
0.55
0.69
5
0
14
10
10
14
30
7
12
14
10
10
14
31
171
172
14
32
11
8
10
13
11
0
8
25
11
10
0
0
0
0
0
6
24
13
12
8
23
0
15
36
11
8
10
13
11
4
8
26
11
10
14
8
16
5
10
6
24
15
12
10
24
9
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
85774
86819
86829
87309
88213
89355
90075
99248
99248
85774
19.20
0.15
86819
17.40
0.15
86829
15.90
0.15
87309
17.60
0.15
88213
19.70
0.15
89355
15.60
0.15
90075
15.20
0.15
99248
16.30
0.15
99248
16.30
0.15
137099
D7099
18.20
0.15
138127
D8127
17.10
0.15
138947
D8947
18.70
0.15
142781
E2781
16.10
0.15
142781
E2781
16.10
0.15
142781
E2781
16.10
0.15
143624
E3624
15.90
0.15
143624
E3624
15.90
0.15
154276
F4276
17.60
0.15
159454
F9454
17.90
0.15
159560
F9560
17.00
0.15
159560
F9560
17.00
0.15
159857
F9857
15.40
0.15
162058
G2058
17.80
0.15
162058
G2058
17.80
0.15
162080
G2080
19.80
0.15
162080
G2080
19.80
0.15
162116
G2116
19.30
0.15
162567
G2567
19.90
0.15
162741
G2741
17.30
0.15
162980
G2980
16.70
0.15
163818
G3818
18.40
0.15
172034
H2034
17.80
0.15
0.94 ± 0.01
0.80 ± 0.27
1.43 ± 0.05
0.57 ± 0.16
0.91 ± 0.42
2.04 ± 0.05
2.23 ± 0.08
1.12 ± 0.04
1.14 ± 0.37
0.56 ± 0.02
0.75 ± 0.02
0.45 ± 0.12
1.59 ± 0.05
2.01 ± 0.74
2.03 ± 0.77
2.14 ± 0.04
2.23 ± 1.08
1.06 ± 0.35
0.58 ± 0.02
1.10 ± 0.47
1.16 ± 0.30
3.07 ± 1.32
0.85 ± 0.01
0.85 ± 0.28
0.78 ± 0.06
0.82 ± 0.33
0.54 ± 0.17
0.33 ± 0.01
3.95 ± 0.04
0.79 ± 0.04
0.39 ± 0.02
0.63 ± 0.02
1.40 ± 0.00
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.47
1.40 ± 0.51
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.48
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.40
1.40 ± 0.00
1.40 ± 0.53
1.40 ± 0.43
1.40 ± 0.00
1.40 ± 0.54
1.40 ± 0.39
1.40 ± 0.45
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.11
1.40 ± 0.49
1.40 ± 0.40
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
0.04 ± 0.01
0.30 ± 0.22
0.37 ± 0.05
0.50 ± 0.23
0.03 ± 0.03
0.25 ± 0.03
0.29 ± 0.04
0.43 ± 0.06
0.41 ± 0.28
0.29 ± 0.04
0.45 ± 0.06
0.29 ± 0.28
0.25 ± 0.04
0.16 ± 0.15
0.16 ± 0.09
0.17 ± 0.03
0.15 ± 0.17
0.14 ± 0.17
0.37 ± 0.04
0.24 ± 0.23
0.21 ± 0.21
0.13 ± 0.16
0.19 ± 0.02
0.19 ± 0.14
0.04 ± 0.01
0.03 ± 0.07
0.12 ± 0.08
0.17 ± 0.03
0.01 ± 0.00
0.66 ± 0.13
0.52 ± 0.06
0.34 ± 0.05
18
0.90
0.79
0.33
0.67
0.66
1.19
0.73
0.29
0.48
0.65
0.17
0.46
0.15
0.85
0.45
0.32
0.82
0.29
0.30
1.16
0.53
0.34
0.34
0.87
1.39
0.99
0.47
0.20
0.22
0.40
0.33
1.05
11
0
14
0
0
30
12
7
0
6
7
0
14
0
0
9
0
0
6
0
0
0
26
0
4
0
0
6
6
8
7
11
7
14
10
6
31
12
8
8
6
7
9
14
15
9
9
8
5
6
87
13
5
27
31
4
13
7
6
6
8
7
16
16
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
190166
J0166
17.10
0.15
190166
J0166
17.10
0.15
209924
K9924
16.10
0.15
211871
L1871
18.80
0.15
214088
L4088
15.20
0.15
215588
L5588
19.50
0.15
215757
L5757
17.70
0.15
235086
N5086
17.50
0.15
235086
N5086
17.50
0.15
235086
N5086
17.50
0.15
242450
O2450
14.70
0.15
242450
O2450
14.70
0.15
250620
P0620
18.00
0.15
267337
Q7337
18.00
0.15
269690
Q9690
18.40
0.15
271480
R1480
17.50
0.15
274138
R4138
17.80
0.15
275976
R5976
16.30
0.15
275976
R5976
16.30
0.15
276274
R6274
17.20
0.15
276468
R6468
17.90
0.15
285944
S5944
16.50
0.15
285944
S5944
16.50
0.15
297418
T7418
18.60
0.15
299582
T9582
18.00
0.15
303174
U3174
16.70
0.15
304330
U4330
18.90
0.15
304330
U4330
18.90
0.15
322763 W2763
16.90
0.15
326388 W6388
18.20
0.15
334673
X4673
17.90
0.15
349219
Y9219
18.20
0.15
1.01 ± 0.03
1.05 ± 0.02
1.86 ± 0.71
0.41 ± 0.01
2.42 ± 0.06
0.49 ± 0.16
0.78 ± 0.27
1.02 ± 0.40
1.02 ± 0.32
1.08 ± 0.33
2.54 ± 0.10
2.91 ± 0.08
0.65 ± 0.14
0.44 ± 0.10
0.89 ± 0.43
0.71 ± 0.22
0.75 ± 0.02
1.86 ± 0.04
2.38 ± 0.03
1.53 ± 0.71
1.03 ± 0.37
1.04 ± 0.04
1.40 ± 0.43
0.41 ± 0.02
0.62 ± 0.02
1.50 ± 0.03
0.61 ± 0.01
0.78 ± 0.01
1.25 ± 0.03
1.26 ± 0.57
0.57 ± 0.22
0.58 ± 0.15
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.48
1.40 ± 0.51
1.40 ± 0.38
1.40 ± 0.38
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.33
1.40 ± 0.43
1.40 ± 0.59
1.40 ± 0.48
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.42
1.40 ± 0.00
1.40 ± 0.41
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.52
1.40 ± 0.60
1.40 ± 0.41
0.25 ± 0.04
0.23 ± 0.03
0.19 ± 0.12
0.32 ± 0.05
0.25 ± 0.03
0.12 ± 0.12
0.24 ± 0.17
0.17 ± 0.18
0.17 ± 0.11
0.15 ± 0.12
0.36 ± 0.13
0.27 ± 0.04
0.26 ± 0.13
0.58 ± 0.25
0.10 ± 0.11
0.35 ± 0.22
0.24 ± 0.03
0.15 ± 0.03
0.09 ± 0.01
0.10 ± 0.17
0.11 ± 0.14
0.41 ± 0.03
0.23 ± 0.19
0.39 ± 0.05
0.29 ± 0.03
0.16 ± 0.03
0.13 ± 0.02
0.08 ± 0.01
0.20 ± 0.04
0.06 ± 0.12
0.38 ± 0.25
0.27 ± 0.23
19
0.92
0.68
0.40
0.28
0.63
0.57
0.47
1.04
1.64
0.85
0.41
0.83
0.29
0.21
0.31
0.82
0.48
1.01
1.11
0.91
0.36
0.16
0.51
0.93
0.31
0.65
0.13
0.23
0.27
0.33
0.67
0.58
6
12
0
5
8
0
0
0
0
0
11
13
0
0
0
0
7
5
15
0
0
10
0
5
7
21
11
12
12
0
0
0
7
12
7
7
8
5
11
60
32
29
11
14
4
4
7
6
7
5
16
5
5
10
29
5
7
23
11
12
13
8
11
14
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
363505
a3505
18.10
0.15
368184
a8184
19.50
0.15
369264
a9264
16.30
0.15
377732
b7732
17.00
0.15
377732
b7732
17.00
0.15
381677
c1677
18.40
0.15
381677
c1677
18.40
0.15
387733
c7733
18.90
0.15
387733
c7733
18.90
0.15
387746
c7746
20.00
0.15
388838
c8838
19.50
0.15
388838
c8838
19.50
0.15
389694
c9694
18.20
0.15
391211
d1211
18.50
0.15
393359
d3359
19.20
0.15
393569
d3569
20.20
0.15
399433
d9433
18.60
0.15
399433
d9433
18.60
0.15
406952
e6952
17.10
0.15
408751
e8751
19.00
0.15
409256
e9256
18.20
0.15
409836
e9836
18.10
0.15
410088
f0088
18.10
0.15
410778
f0778
18.10
0.15
411201
f1201
17.80
0.15
411611
f1611
18.80
0.15
413038
f3038
16.90
0.15
413038
f3038
16.90
0.15
413192
f3192
16.80
0.15
413421
f3421
18.30
0.15
413820
f3820
19.80
0.15
414286
f4286
18.60
0.15
1.90 ± 0.05
0.38 ± 0.12
1.51 ± 0.47
0.95 ± 0.03
0.99 ± 0.03
0.47 ± 0.01
0.44 ± 0.16
0.34 ± 0.01
0.32 ± 0.09
0.37 ± 0.01
0.36 ± 0.01
0.38 ± 0.01
0.45 ± 0.02
0.41 ± 0.09
0.77 ± 0.33
0.55 ± 0.01
1.34 ± 0.56
1.76 ± 0.89
0.77 ± 0.21
0.40 ± 0.01
1.89 ± 0.68
0.55 ± 0.19
1.03 ± 0.01
1.46 ± 0.57
0.66 ± 0.01
0.36 ± 0.10
1.24 ± 0.03
1.01 ± 0.04
3.96 ± 1.84
1.90 ± 0.78
0.66 ± 0.26
0.37 ± 0.08
1.40 ± 0.03
1.40 ± 0.46
1.40 ± 0.42
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.54
1.40 ± 0.00
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.38
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.49
1.40 ± 0.53
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.40
1.40 ± 0.49
1.40 ± 0.00
1.40 ± 0.41
1.40 ± 0.00
1.40 ± 0.43
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.47
1.40 ± 0.41
1.40 ± 0.46
1.40 ± 0.38
0.03 ± 0.01
0.19 ± 0.19
0.23 ± 0.20
0.31 ± 0.05
0.29 ± 0.03
0.35 ± 0.05
0.40 ± 0.21
0.41 ± 0.06
0.47 ± 0.25
0.13 ± 0.02
0.21 ± 0.04
0.20 ± 0.02
0.46 ± 0.06
0.42 ± 0.23
0.06 ± 0.11
0.05 ± 0.01
0.04 ± 0.09
0.02 ± 0.05
0.43 ± 0.23
0.28 ± 0.03
0.03 ± 0.04
0.33 ± 0.25
0.10 ± 0.02
0.05 ± 0.03
0.31 ± 0.05
0.41 ± 0.21
0.20 ± 0.04
0.30 ± 0.04
0.02 ± 0.05
0.02 ± 0.02
0.05 ± 0.04
0.47 ± 0.19
20
0.66
0.54
0.65
0.62
0.16
0.92
0.47
0.22
0.37
0.23
0.61
0.24
0.22
0.85
0.51
0.22
0.24
0.16
0.64
0.84
0.84
1.78
0.14
0.38
1.45
0.81
1.19
1.79
0.62
1.39
0.97
0.54
12
0
0
5
5
19
0
11
0
5
18
12
4
0
0
13
0
0
0
68
0
0
9
0
12
0
22
23
0
0
0
0
12
25
7
5
5
19
5
11
5
6
18
12
5
17
30
14
10
9
5
69
4
14
10
6
15
31
23
25
18
23
36
27
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
414286
418797
418929
419624
419624
419880
f4286
18.60
0.15
f8797
19.40
0.15
f8929
17.00
0.15
f9624
20.50
0.15
f9624
20.50
0.15
f9880
19.60
0.15
2000 AG205 K00AK5G 19.70
0.15
2002 XS40
K02X40S
20.10
0.15
2003 CC11
K03C11C 19.10
0.15
2003 SS214
K03SL4S
20.10
0.15
2004 BZ74
K04B74Z
18.10
0.15
2004 MX2
K04M02X 19.30
0.15
2004 TG10
K04T10G 19.40
0.15
2005 LS3
K05L03S
19.50
0.15
2006 BB27
K06B27B 20.00
0.15
2007 BG
K07B00G 19.50
0.15
2007 RU10
K07R10U 19.10
0.15
2008 QS11
K08Q11S
19.90
0.15
2009 ND1
K09N01D 17.10
0.15
2010 OQ1
K10O01Q 19.00
0.15
2011 CQ4
K11C04Q 18.40
0.15
2012 DN
K12D00N 18.10
0.15
2013 PX6
K13P06X 18.40
0.15
2013 WT44 K13W44T 19.30
0.15
2013 WU44 K13W44U 21.00
0.15
2013 YZ13
K13Y13Z
19.60
0.15
2013 YP139 K13YD9P
21.60
0.15
2014 AA33
K14A33A 19.30
0.15
2014 AQ46
K14A46Q 20.10
0.15
2014 AA53
K14A53A 19.80
0.15
2014 BG60
K14B60G 20.10
0.15
2014 BE63
K14B63E
23.20
0.15
0.40 ± 0.09
0.70 ± 0.29
1.43 ± 0.02
0.34 ± 0.14
0.36 ± 0.13
0.98 ± 0.06
0.95 ± 0.01
0.76 ± 0.03
1.13 ± 0.51
0.86 ± 0.25
0.96 ± 0.02
1.26 ± 0.08
1.32 ± 0.61
0.38 ± 0.10
0.22 ± 0.05
0.31 ± 0.11
0.92 ± 0.37
0.45 ± 0.01
2.50 ± 0.95
0.54 ± 0.21
0.66 ± 0.02
2.77 ± 1.05
1.65 ± 0.03
0.65 ± 0.01
0.29 ± 0.13
0.31 ± 0.10
0.40 ± 0.03
0.79 ± 0.04
0.59 ± 0.29
0.70 ± 0.27
0.67 ± 0.25
0.36 ± 0.13
0.40 ± 0.24
0.06 ± 0.07
0.14 ± 0.03
0.09 ± 0.17
0.09 ± 0.14
0.03 ± 0.01
0.03 ± 0.00
0.03 ± 0.00
0.03 ± 0.10
0.02 ± 0.02
0.11 ± 0.02
0.02 ± 0.00
0.02 ± 0.04
0.19 ± 0.12
0.38 ± 0.21
0.24 ± 0.19
0.05 ± 0.06
0.09 ± 0.01
0.04 ± 0.04
0.15 ± 0.14
0.18 ± 0.02
0.01 ± 0.03
0.03 ± 0.00
0.08 ± 0.02
0.09 ± 0.19
0.27 ± 0.19
0.03 ± 0.01
0.05 ± 0.01
0.05 ± 0.11
0.04 ± 0.06
0.04 ± 0.08
0.01 ± 0.00
1.40 ± 0.40
1.40 ± 0.50
1.40 ± 0.00
1.40 ± 0.50
1.40 ± 0.46
1.40 ± 0.08
1.40 ± 0.00
1.40 ± 0.05
1.40 ± 0.53
1.40 ± 0.35
1.40 ± 0.00
1.40 ± 0.09
1.40 ± 0.51
1.40 ± 0.38
1.40 ± 0.38
1.40 ± 0.51
1.40 ± 0.47
1.40 ± 0.00
1.40 ± 0.39
1.40 ± 0.51
1.40 ± 0.00
1.40 ± 0.38
1.40 ± 0.02
1.40 ± 0.00
1.40 ± 0.61
1.40 ± 0.46
1.09 ± 0.07
1.40 ± 0.06
1.40 ± 0.60
1.40 ± 0.47
1.40 ± 0.46
1.40 ± 0.46
0.71
0.32
0.54
0.57
0.46
0.20
0.79
0.18
0.49
0.75
0.65
0.35
0.64
0.64
0.98
0.38
0.31
0.33
0.70
0.47
0.29
0.42
0.19
0.31
0.26
0.07
0.25
0.19
0.47
0.50
1.30
0.42
0
0
48
0
0
6
12
14
0
0
4
9
0
0
0
3
0
9
0
0
5
0
9
6
0
0
6
4
0
0
0
0
29
7
49
18
6
6
14
14
16
16
5
9
8
7
5
5
9
11
11
8
7
7
10
6
8
6
6
4
17
13
163
5
21
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
2014 CY4
K14C04Y 21.10
0.15
2014 DC10
K14D10C 20.10
0.15
2014 ED
K14E00D 19.30
0.15
2014 EN45
K14E45N 21.20
0.15
2014 EZ48
K14E48Z
18.80
0.15
2014 EZ48
K14E48Z
18.80
0.15
2014 EQ49
K14E49Q 21.80
0.15
2014 ER49
K14E49R 18.60
0.15
2014 HE3
K14H03E
19.90
0.15
2014 HQ124 K14HC4Q 18.90
0.15
2014 HF177 K14HH7F
19.70
0.15
2014 JL25
K14J25L
23.00
0.15
2014 JH57
K14J57H 16.60
0.15
2014 JH57
K14J57H 16.60
0.15
2014 JN57
K14J57N 20.70
0.15
2014 KX99
K14K99X 18.20
0.15
2014 LQ25
K14L25Q 20.00
0.15
2014 LR26
K14L26R 18.50
0.15
2014 MQ18 K14M18Q 15.60
0.15
2014 NB39
K14N39B 19.50
0.15
2014 NE52
K14N52E
17.90
0.15
2014 NC64
K14N64C 20.50
0.15
2014 NM64 K14N64M 22.60
0.15
2014 OY1
K14O01Y 19.10
0.15
2014 OZ1
K14O01Z
21.00
0.15
2014 PC68
K14P68C 20.40
0.15
2014 PF68
K14P68F
18.20
0.15
2014 QK433 K14Qh3K 18.30
0.15
2014 RH12
K14R12H 23.50
0.15
2014 RL12
K14R12L
17.90
0.15
2014 RL12
K14R12L
17.90
0.15
2014 SR339
K14SX9R 18.60
0.15
0.57 ± 0.25
0.89 ± 0.01
0.49 ± 0.13
0.37 ± 0.13
0.45 ± 0.01
0.44 ± 0.11
0.38 ± 0.13
0.46 ± 0.15
0.56 ± 0.15
0.41 ± 0.17
0.25 ± 0.01
0.23 ± 0.06
4.61 ± 0.03
6.79 ± 3.81
0.27 ± 0.10
1.72 ± 0.68
0.94 ± 0.32
2.08 ± 0.90
5.27 ± 3.50
1.08 ± 0.15
0.70 ± 0.22
0.50 ± 0.19
0.33 ± 0.12
0.60 ± 0.21
0.73 ± 0.29
0.56 ± 0.20
3.33 ± 2.06
1.78 ± 0.75
0.09 ± 0.04
0.69 ± 0.02
0.61 ± 0.17
0.97 ± 0.37
0.02 ± 0.04
0.02 ± 0.00
0.14 ± 0.14
0.04 ± 0.01
0.26 ± 0.04
0.27 ± 0.21
0.02 ± 0.03
0.30 ± 0.26
0.06 ± 0.04
0.29 ± 0.22
0.36 ± 0.06
0.02 ± 0.03
0.02 ± 0.00
0.01 ± 0.03
0.12 ± 0.10
0.03 ± 0.05
0.02 ± 0.01
0.02 ± 0.03
0.04 ± 0.07
0.02 ± 0.02
0.25 ± 0.27
0.04 ± 0.02
0.01 ± 0.02
0.11 ± 0.09
0.01 ± 0.03
0.04 ± 0.04
0.01 ± 0.01
0.03 ± 0.06
0.09 ± 0.11
0.25 ± 0.03
0.33 ± 0.19
0.07 ± 0.07
1.40 ± 0.52
1.40 ± 0.00
1.40 ± 0.39
0.75 ± 0.24
1.40 ± 0.00
1.40 ± 0.38
1.40 ± 0.42
1.40 ± 0.49
1.40 ± 0.34
1.40 ± 0.57
1.40 ± 0.00
1.40 ± 0.34
1.40 ± 0.00
1.40 ± 0.47
1.40 ± 0.47
1.40 ± 0.46
1.40 ± 0.37
1.40 ± 0.46
1.40 ± 0.52
1.40 ± 0.18
1.40 ± 0.47
0.82 ± 0.29
1.40 ± 0.44
1.40 ± 0.43
1.40 ± 0.49
1.40 ± 0.43
1.20 ± 0.48
1.40 ± 0.47
1.40 ± 0.54
1.40 ± 0.00
1.40 ± 0.42
1.40 ± 0.46
0.35
0.90
0.57
0.16
1.10
0.47
0.42
0.51
0.18
0.80
0.39
0.68
0.11
0.30
0.69
0.43
0.48
0.65
0.54
0.08
0.66
0.64
0.82
0.30
0.38
0.39
0.60
0.79
0.75
0.31
0.83
0.69
0
9
0
12
5
0
0
0
0
0
10
0
6
0
0
0
0
0
0
7
0
5
0
0
0
0
0
0
0
5
0
0
5
10
6
12
6
6
5
9
5
10
12
5
6
5
4
9
5
6
8
7
9
6
25
6
21
8
12
10
10
5
6
13
22
Table 2—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
2014 TW57 K14T57W 20.10
0.15
2014 TF64
K14T64F
20.10
0.15
2014 TJ64
K14T64J
21.30
0.15
2014 TJ64
K14T64J
21.30
0.15
2014 UG176 K14UH6G 21.50
0.15
2014 US192
K14UJ2S
18.70
0.15
2014 UF206 K14UK6F
18.80
0.15
2014 UH210
K14UL0H 21.10
0.15
2014 VP35
K14V35P
22.70
0.15
2014 WJ70
K14W70J
17.60
0.15
2014 XQ7
K14X07Q 20.60
0.15
2014 XX7
K14X07X 19.80
0.15
2014 XX31
K14X31X 17.60
0.15
0.47 ± 0.01
0.70 ± 0.20
0.52 ± 0.20
0.52 ± 0.23
0.42 ± 0.12
0.87 ± 0.01
1.63 ± 0.79
0.40 ± 0.16
0.12 ± 0.05
2.92 ± 1.21
0.65 ± 0.29
1.20 ± 0.38
1.35 ± 0.49
0.07 ± 0.02
0.03 ± 0.03
0.02 ± 0.02
0.02 ± 0.03
0.03 ± 0.03
0.08 ± 0.01
0.02 ± 0.04
0.04 ± 0.06
0.10 ± 0.10
0.02 ± 0.04
0.02 ± 0.05
0.01 ± 0.02
0.09 ± 0.15
1.40 ± 0.00
1.40 ± 0.35
1.40 ± 0.47
1.40 ± 0.54
1.40 ± 0.39
1.40 ± 0.00
1.40 ± 0.49
1.40 ± 0.47
1.40 ± 0.53
1.40 ± 0.44
1.40 ± 0.55
1.40 ± 0.36
1.40 ± 0.43
0.76
0.33
0.46
0.55
0.17
0.25
0.62
0.76
0.36
0.62
0.83
0.43
0.42
4
0
0
0
0
5
0
0
0
0
0
0
0
6
5
31
14
8
5
17
5
6
27
8
6
8
23
Table 3. Measured diameters (d) and albedos (pV ) of near-Earth asteroids. Objects in
this table have previously reported measurements by the NEOWISE team (Mainzer et al.
2011b, 2012). Previous measurements use detections in 12 µm and 22 µm bands, and
therefore are better constrained. Magnitude H, slope parameter G, and beaming η used
are given. The numbers of observations used in the 3.4 µm (nW 1) and 4.6 µm (nW 2)
wavelengths are also reported, along with the amplitude of the 4.6 µm light curve (W2
amp.).
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
2102
2102
2102
3554
4183
4183
6050
25916
27346
40263
40267
85628
90367
90367
137062
138847
162181
162483
162566
163691
243566
262623
276049
277616
395207
395207
397237
02102
16.00
0.15
02102
16.00
0.15
02102
16.00
0.15
03554
15.82
0.15
04183
14.40
0.15
04183
14.40
0.15
06050
14.80
0.15
25916
13.60
0.15
27346
15.90
0.15
40263
17.70
0.15
40267
15.40
0.15
85628
17.00
0.15
90367
17.70
0.15
90367
17.70
0.15
D7062
16.60
0.15
D8847
16.90
0.15
G2181
18.20
0.15
G2483
17.50
0.15
G2566
15.70
0.15
G3691
17.00
0.15
O3566
17.40
0.15
Q2623
18.50
0.15
R6049
16.80
0.15
R7616
17.40
0.15
d5207
19.60
0.15
d5207
19.60
0.15
d7237
16.70
0.15
1998 SB15
J98S15B 20.90
0.15
2009 UX17 K09U17X 21.50
0.15
2010 LF86
K10L86F
17.30
0.15
2010 LO97 K10L97O 18.70
0.15
2010 NG3 K10N03G 17.20
0.15
1.68 ± 0.05
1.65 ± 0.05
1.69 ± 0.06
1.56 ± 0.07
2.94 ± 0.12
3.54 ± 0.12
2.88 ± 0.07
5.96 ± 0.13
1.80 ± 0.07
0.92 ± 0.35
2.39 ± 0.09
0.78 ± 0.03
1.76 ± 0.79
2.00 ± 0.89
0.99 ± 0.06
0.94 ± 0.28
0.73 ± 0.02
0.69 ± 0.20
6.00 ± 2.42
3.06 ± 1.55
0.88 ± 0.02
0.49 ± 0.15
4.03 ± 1.85
1.28 ± 0.01
0.60 ± 0.20
0.73 ± 0.30
1.73 ± 0.66
0.36 ± 0.12
0.39 ± 0.13
2.30 ± 0.89
1.40 ± 0.57
1.45 ± 0.02
0.25 ± 0.04
0.26 ± 0.04
0.25 ± 0.03
0.34 ± 0.06
0.36 ± 0.06
0.24 ± 0.04
0.26 ± 0.04
0.18 ± 0.03
0.24 ± 0.04
0.17 ± 0.18
0.21 ± 0.04
0.46 ± 0.08
0.05 ± 0.13
0.04 ± 0.03
0.41 ± 0.05
0.35 ± 0.19
0.17 ± 0.03
0.37 ± 0.21
0.03 ± 0.04
0.03 ± 0.06
0.25 ± 0.04
0.29 ± 0.18
0.02 ± 0.04
0.12 ± 0.02
0.07 ± 0.03
0.05 ± 0.10
0.12 ± 0.16
0.06 ± 0.09
0.03 ± 0.03
0.04 ± 0.04
0.03 ± 0.06
0.11 ± 0.02
24
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.48
1.40 ± 0.00
1.40 ± 0.00
1.40 ± 0.51
1.40 ± 0.46
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.40
1.40 ± 0.54
1.40 ± 0.00
1.40 ± 0.44
1.40 ± 0.44
1.40 ± 0.00
1.40 ± 0.40
1.40 ± 0.49
1.40 ± 0.46
1.40 ± 0.44
1.40 ± 0.40
1.40 ± 0.41
1.40 ± 0.47
1.40 ± 0.00
0.23
0.18
0.67
0.49
1.04
0.62
1.51
0.63
0.43
0.71
1.06
0.64
0.54
0.49
0.89
1.01
0.32
0.62
1.02
0.30
0.29
0.48
0.54
0.28
0.32
0.50
0.40
0.66
0.86
0.21
0.57
0.64
13
5
8
19
12
17
57
24
9
0
4
7
0
0
6
0
25
0
0
0
11
0
0
4
0
0
0
0
0
0
0
17
13
5
9
20
12
18
57
29
9
14
4
10
12
13
6
26
25
9
24
5
11
4
6
4
8
19
4
11
15
7
15
17
Table 3—Continued
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
2010 NG3 K10N03G 17.20
0.15
2014 HJ129 K14HC9J
21.10
0.15
1.74 ± 0.94
0.59 ± 0.21
0.08 ± 0.18
0.02 ± 0.02
1.40 ± 0.59
1.40 ± 0.42
0.80
0.50
0
0
17
9
25
Table 4. Measured diameters (d) and albedos (pV ) of MBAs and Mars crossers. Objects
in this table do not have previously published diameters and albedos by the NEOWISE
team. Beaming η, H, G, the amplitude of the 4.6 µm light curve (W2 amp.), and the
numbers of observations used in the 3.4 µm (nW 1) and 4.6 µm (nW 2) wavelengths are also
reported. For a small (< 1%) fraction of objects, diameter fits could not reproduce optical
magnitudes for a realistic range of albedos. This may be due to a large light curve
amplitude, uncertainty in G slope values used to derive H magnitudes, or other reasons
noted in Mainzer et al. (2011b); Masiero et al. (2011); Mainzer et al. (2012); Masiero et al.
(2012). These objects are marked with a † in the name column. Objects without reported
albedos did not have measured H values, see text for details. Only the first 15 lines are
shown; the remainder are available in electronic format through the journal website.
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
21
65
69
74
74
140
144
147
147
160
212
212
253
284
00021
7.35
0.11
00065
6.62
0.01
00069
7.05
0.19
00074
8.66
0.15
00074
8.66
0.15
00140
8.34
0.15
00144
7.91
0.17
00147
8.70
0.15
00147
8.70
0.15
00160
9.08
0.15
00212
8.28
0.15
00212
8.28
0.15
00253
10.30
0.15
00284
10.05
0.11
99.47 ± 27.12
276.58 ± 74.49
131.07 ± 32.19
111.87 ± 46.38
105.13 ± 29.95
82.63 ± 20.19
131.36 ± 33.30
144.68 ± 47.63
119.59 ± 37.39
69.62 ± 13.23
132.58 ± 48.48
129.09 ± 40.48
50.35 ± 17.16
54.47 ± 20.59
0.16 ± 0.12
0.06 ± 0.04
0.19 ± 0.07
0.04 ± 0.03
0.05 ± 0.02
0.09 ± 0.07
0.05 ± 0.01
0.03 ± 0.02
0.04 ± 0.02
0.07 ± 0.04
0.05 ± 0.03
0.05 ± 0.04
0.04 ± 0.02
0.04 ± 0.03
0.95 ± 0.19
0.95 ± 0.17
0.95 ± 0.18
0.95 ± 0.23
0.95 ± 0.16
0.95 ± 0.18
0.95 ± 0.17
0.95 ± 0.19
0.95 ± 0.18
0.95 ± 0.14
0.95 ± 0.20
0.95 ± 0.19
0.95 ± 0.24
0.95 ± 0.23
0.27
0.09
0.09
0.26
0.24
0.37
0.31
0.11
0.20
0.58
0.16
0.17
0.43
0.21
10
8
14
3
9
7
10
6
9
20
5
7
16
11
10
10
14
4
9
7
10
6
9
21
5
7
16
11
26
27
Table 5. Measured diameters (d) and albedos (pV ) of MBA and Mars crossing asteroids.
Objects in this table have previously reported measurements by the NEOWISE team
(Masiero et al. 2011, 2012). Previous measurements use detections in 12 µm and 22 µm
bands, and therefore are better constrained. H, G, beaming η, the amplitude of the 4.6 µm
light curve (W2 amp.), and the numbers of observations used in the 3.4 µm (nW 1) and 4.6
µm (nW 2) wavelengths are also reported. For a small (< 1%) fraction of objects, diameter
fits could not reproduce optical magnitudes for a realistic range of albedos. This may be
due to a large light curve amplitude, uncertainty in G slope values used to derive H
magnitudes, or other reasons noted in Mainzer et al. (2011b); Masiero et al. (2011);
Mainzer et al. (2012); Masiero et al. (2012). These objects are marked with a † in the name
column. Only the first 15 lines are shown; the remainder are available in electronic format
through the journal website.
Name
Packed
H
G
d (km)
pV
η
W2 amp.
nW 1
nW 2
13
24
30
33
34
35
36
36
38
40
40
41
45
47†
48†
00013
6.74
0.15
00024
7.08
0.19
00030
7.57
0.15
00033
8.55
0.33
00034
8.51
0.15
00035
8.50
0.15
00036
8.46
0.15
00036
8.46
0.15
00038
8.32
0.15
00040
7.00
0.15
00040
7.00
0.15
00041
7.12
0.10
00045
7.46
0.07
00047
7.84
0.16
00048
6.90
0.15
202.64 ± 50.08
151.82 ± 49.32
93.51 ± 23.53
48.78 ± 9.98
114.12 ± 43.76
143.02 ± 55.51
102.44 ± 31.81
92.34 ± 39.98
114.16 ± 28.20
95.55 ± 17.94
107.07 ± 19.23
198.74 ± 61.71
181.92 ± 59.39
107.18 ± 33.79
165.38 ± 41.80
0.06 ± 0.03
0.08 ± 0.04
0.26 ± 0.15
0.25 ± 0.13
0.04 ± 0.02
0.03 ± 0.03
0.05 ± 0.02
0.06 ± 0.05
0.05 ± 0.01
0.32 ± 0.09
0.29 ± 0.10
0.05 ± 0.07
0.05 ± 0.03
0.07 ± 0.03
0.06 ± 0.03
0.95 ± 0.16
0.95 ± 0.20
0.95 ± 0.21
0.95 ± 0.19
0.95 ± 0.24
0.95 ± 0.21
0.95 ± 0.19
0.95 ± 0.25
0.95 ± 0.16
0.95 ± 0.16
0.95 ± 0.16
0.95 ± 0.20
0.95 ± 0.21
0.95 ± 0.20
0.95 ± 0.15
0.32
0.20
0.39
0.41
0.17
0.34
0.21
0.30
0.13
0.26
0.25
0.13
0.14
0.28
0.42
9
15
13
15
21
11
10
6
14
12
23
19
11
14
11
9
15
13
15
22
11
10
6
14
12
23
19
11
14
11
28
29
Results were compared to previous work by the NEOWISE team (Mainzer et al.
2011b; Masiero et al. 2011; Mainzer et al. 2012; Masiero et al. 2012). Figure 3 shows the
comparison between diameter and albedo measurements of MBAs. As observed in Masiero
et al. (2011), asteroids in the Main Belt group into bright and dark types, with a greater
fraction of bright objects found in the inner regions of the belt. Objects that were also
modeled with the thermophysical model of Wright (2007) are given in Table 6.
When possible, derived diameters were compared to diameter measurements made
from radar data. Radar-derived diameters are ideal for this purpose, as they are derived
via an independent method (Benner et al. 2015). This comparison is shown in Figure 4.
Although the histograms in the figure are not perfectly Gaussian, a best-fit Gaussian to
their forms gives fitted σ values, which indicates a 14% relative accuracy in diameter, and
a 29% relative accuracy in albedo. These values are consistent with previous NEOWISE
3-band data results (Mainzer et al. 2012; Masiero et al. 2012). From this comparison to
radar-derived diameters and previous work, we conclude that diameters are determined to
an accuracy of ∼ 20% or better. If good-quality H magnitudes are available, albedos can be
determined to within ∼ 40% or better.
Roughly 3% of objects in this work have significantly different derived diameters
than previously published NEOWISE values. It is possible that some of these objects
are elongated. NEOWISE collects a sparsely-sampled lightcurve for each object, and for
Table 6. Measured diameters and albedos for three objects using the model of Wright
(2007).
Name
D (km)
pV
68267
138127
285944
0.89 ± 0.27
0.94 ± 0.15
1.37 ± 0.23
0.38 ± 0.32
0.35 ± 0.08
0.34 ± 0.08
30
Fig. 3.— Histogram of MBA diameters (top) and albedos (bottom) measured in this work
(blue), and values for the same objects measured by the NEOWISE team previously (green).
The two albedo peaks are due to the predominance of bright S-type (pV = 0.25) and dark
C-type (pV = 0.06) objects in the Main Belt.
example, it is possible that the prime mission happened to observe these objects in a more
edge-on shape, whereas the reactivation observations tended to observe a wider side of the
object. Alternatively, changes in viewing geometry between epochs could result in different
31
diameter measurements; a pole-on viewing geometry could have a larger cross section than
a geometry aligned with the plane of the equator.
For a small (< 1%) fraction of objects, diameter fits could not reproduce optical
magnitudes for a realistic range of albedos. This may be due to a large light curve amplitude
(see column W2 amp. for the amplitude of the 3.4 µm band light curve, though note that
this is a sparsely sampled light curve), uncertainty in G slope values used to derive H
magnitudes, or other reasons noted in Mainzer et al. (2011b); Masiero et al. (2011); Mainzer
et al. (2012); Masiero et al. (2012). Poor-quality H values can drive albedo fits to extremes;
therefore very low (∼ 0.01) measurements may be signs of this phenomenon.
We have plotted the diameters and albedos of NEOWISE Year One Reactivation
discoveries, along with all NEAs detected by NEOWISE (Figure 5). The trend observed in
Mainzer et al. (2014) is also present here: NEOWISE tends to discover darker NEAs than
optical surveys. This is a direct consequence of the infrared wavelengths that NEOWISE
employs.
3.1. NHATS Targets
Five objects in this paper meet the NEO Human Space Flight Accessible Targets Study
qualifications (Barbee et al. 2013). These objects are listed in Table 7. If an object was
observed over multiple epochs, values of d and pV in this table are the averages of the values
and associated errors derived at each of those epochs. Asteroid 419624 was discovered in
2010 by NEOWISE.
32
4. Conclusion
We present preliminary diameters and albedos for 7,959 asteroids observed in the first
year of the NEOWISE Reactivation mission. Five of these objects are NHATS targets.
Future work by the NEOWISE team includes preliminary characterization results from the
continuing mission.
Uncertainties on d and pV are consistent with the errors measured during the initial
post-cryo mission. NEOWISE is expected to maintain this pace of detection and NEO
discovery for the extent of its mission, currently expected to run through 2017. These
results demonstrate the power of infrared survey telescopes to characterize basic physical
parameters for large numbers of small bodies.
5. Acknowledgments
CRN was partially supported by an appointment to the NASA Postdoctoral Program
at the Jet Propulsion Laboratory (JPL), administered by Oak Ridge Associated Universities
through a contract with NASA. This publication makes use of data products from the
Wide-field Infrared Survey Explorer, which is a joint project of the University of California,
Los Angeles, and JPL/California Institute of Technology, funded by NASA. This publication
also makes use of data products from NEOWISE, which is a project of the JPL/California
Institute of Technology, funded by the Planetary Science Division of NASA. This research
has made use of the NASA/IPAC Infrared Science Archive. The JPL High-Performance
Computing Facility used for our simulations is supported by the JPL Office of the CIO.
This project used data obtained with the Dark Energy Camera (DECam), which was
constructed by the Dark Energy Survey (DES) collaboration. Funding for the DES Projects
has been provided by the U.S. Department of Energy, the U.S. National Science Foundation,
33
the Ministry of Science and Education of Spain, the Science and Technology Facilities
Council of the United Kingdom, the Higher Education Funding Council for England, the
National Center for Supercomputing Applications at the University of Illinois at Urbana-
Champaign, the Kavli Institute of Cosmological Physics at the University of Chicago, the
Center for Cosmology and Astro-Particle Physics at the Ohio State University, the Mitchell
Institute for Fundamental Physics and Astronomy at Texas A&M University, Financiadora
de Estudos e Projetos, Funda¸cao Carlos Chagas Filho de Amparo `a Pesquisa do Estado do
Rio de Janeiro, Conselho Nacional de Desenvolvimento Cient´ıfico e Tecnol´ogico and the
Minist´erio da Ciencia, Tecnologia e Inovacao, the Deutsche Forschungsgemeinschaft, and
the Collaborating Institutions in the Dark Energy Survey. The Collaborating Institutions
are Argonne National Laboratory, the University of California at Santa Cruz, the University
of Cambridge, Centro de Investigaciones En´ergeticas, Medioambientales y Tecnol´ogicas-
Madrid, the University of Chicago, University College London, the DES-Brazil Consortium,
the University of Edinburgh, the Eidgenossische Technische Hochschule (ETH) Zurich,
Fermi National Accelerator Laboratory, the University of Illinois at Urbana-Champaign,
the Institut de Ci`encies de l’Espai (IEEC/CSIC), the Institut de F´ısica d’Altes Energies,
Lawrence Berkeley National Laboratory, the Ludwig-Maximilians Universitat Munchen
and the associated Excellence Cluster Universe, the University of Michigan, the National
Optical Astronomy Observatory, the University of Nottingham, the Ohio State University,
the University of Pennsylvania, the University of Portsmouth, SLAC National Accelerator
Laboratory, Stanford University, the University of Sussex, and Texas A&M University.
This work makes use of observations from the LCOGT network.
Follow-up included observations obtained at the Gemini Observatory, which is operated
by the Association of Universities for Research in Astronomy, Inc., under a cooperative
agreement with the NSF on behalf of the Gemini partnership: the National Science
Foundation (United States), the National Research Council (Canada), CONICYT (Chile),
34
the Australian Research Council (Australia), Ministrio da Cincia, Tecnologia e Inovao
(Brazil) and Ministerio de Ciencia, Tecnologa e Innovacin Productiva (Argentina).
We thank the anonymous referee for their thoughtful and thorough consideration of
our manuscript.
35
Fig. 4.— Top: Comparison of radar-derived diameters and albedos to the values derived in
this paper. Dashed red line is shows a 1:1 relation. Bottom: %∆d (left) and %∆pV (right)
are the fractional differences between the NEOWISE Reactivation radar-derived diameters
and albedos, respectively. Dashed red line is best-fit Gaussian, with the fitted σ given in the
legends.
36
Fig. 5.— NEOWISE detects large NEAs, and discoveries tend to be dark. Cyan circles
are measured diameters and albedos of objects detected in the first year of NEOWISE’s
Reactivation mission; black squares indicate NEAs discovered by NEOWISE. Error bars on
detected objects omitted for clarity.
Table 7. Measured diameters and albedos for objects that meet NHATS criteria. Also
included are the minimum round trip time in days, as determined by the Barbee et al.
(2013) study.
Name
D (km)
pV
Minimum round trip (days)
1943 Anteros
35107
363505
387733
419624
2.30 ± 0.05
1.00 ± 0.15
1.90 ± 0.05
0.33 ± 0.05
0.35 ± 0.13
0.17 ± 0.02
0.34 ± 0.10
0.03 ± 0.01
0.44 ± 0.15
0.09 ± 0.15
354
354
314
354
362
37
REFERENCES
Barbee, B. W., Abell, P. A., Adamo, D. R., et al. 2013, 2013 IAA Planetary Defense
Conference Proceedings
Bauer, J. M., Grav, T., Blauvelt, E., et al. 2013, The Astrophysical Journal, 773, 22
Benner, L., Busch, M. B., Giorgini, J. D., Taylor, P. A., & Margot, J. L. 2015, Asteroids IV
Brown, T. M., Baliber, N., Bianco, F. B., et al. 2013, Publications of the Astronomical
Society of the Pacific, 125, 1031
Bus, S. J. & Binzel, R. P. 2002, Icarus, 158, 146
Carruba, V., Domingos, R. C., Nesvorn´y, D., et al. 2013, Monthly Notices of the Royal
Astronomical Society, 433, 2075
Cruikshank, D. P. 1977, Icarus, 30, 224
Cutri, R. M., Mainzer, A., Conrow, T., et al. 2015, 1, explana-
tory Supplement to the NEOWISE Data Release Products,
http://wise2.ipac.caltech.edu/docs/release/neowise/expsup
Cutri, R. M., Wright, E. L., Conrow, T., et al. 2012, 1,
http://wise2.ipac.caltech.edu/docs/release/allsky/expsup/sec8 1.html
Delb´o, M., Harris, A. W., Binzel, R. P., Pravec, P., & Davies, J. K. 2003, Icarus, 166, 116
Delb´o, M., Walsh, K., Mueller, M., Harris, A. W., & Howell, E. S. 2011, Icarus, 212, 138
Gradie, J. & Tedesco, E. 1982, Science, 216, 1405
Grav, T., Mainzer, A. K., Bauer, J., et al. 2012a, The Astrophysical Journal, 744, 197
—. 2011, The Astrophysical Journal, 742, 40
38
Grav, T., Mainzer, A. K., Bauer, J. M., Masiero, J. R., & Nugent, C. R. 2012b, The
Astrophysical Journal, 759, 49
—. 2012c, The Astrophysical Journal, 759, 49
Hansen, O. L. 1976, The Astronomical Journal, 81, 74
Harris, A. W. 1998, Icarus, 131, 291
Harris, A. W. & Lagerros, J. S. V. 2002, Asteroids III, 205
Lebofsky, L. A., Veeder, G. J., Lebofsky, M. J., & Matson, D. L. 1978, Icarus, 35, 336
Mainzer, A., Bauer, J., Cutri, R. M., et al. 2014, The Astrophysical Journal, 792, 30
Mainzer, A., Bauer, J., Grav, T., et al. 2011a, The Astrophysical Journal, 731, 53
Mainzer, A., Grav, T., Bauer, J., et al. 2011b, The Astrophysical Journal, 743, 156
Mainzer, A., Grav, T., Masiero, J., et al. 2012, The Astrophysical Journal Letters, 760, L12
—. 2011c, The Astronomical Journal, 736
Mainzer, A., Usui, F., & Trilling, D. E. 2015, Asteroids IV
Masiero, J. R., DeMeo, F., Kasuga, T., & Parker, A. H. 2015, Asteroids IV
Masiero, J. R., Grav, T., Mainzer, A. K., et al. 2014, The Astrophysical Journal, 791, 121
Masiero, J. R., Mainzer, A. K., Bauer, J. M., et al. 2013, The Astrophysical Journal, 770, 7
Masiero, J. R., Mainzer, A. K., Grav, T., et al. 2011, The Astrophysical Journal, 741, 68
—. 2012, The Astrophysical Journal Letters, 759, L8
Matter, A., Delb´o, M., Ligori, S., Crouzet, N., & Tanga, P. 2011, Icarus, 215, 47
39
McMillan, R. S. 2007, in IAU Symposium, Vol. 236, IAU Symposium, ed. G. B. Valsecchi,
D. Vokrouhlick´y, & A. Milani, 329–340
Milani, A., Cellino, A., Knezevi´c, Z., et al. 2014, Icarus, 239, 46
Morrison, D. & Lebofsky, L. 1979, Radiometry of asteroids, ed. T. Gehrels, 184–205
Muller, T. G., Miyata, T., Kiss, C., et al. 2013, Astronomy & Astrophysics, 558, A97
Muller, T. G., O’Rourke, L., Barucci, A. M., et al. 2012, Astronomy & Astrophysics, 548,
A36
Ostro, S. J., Hudson, R. S., Benner, L. A. M., et al. 2002, Asteroids III, 151
Pravec, P., Harris, A. W., Kusnir´ak, P., Gal´ad, A., & Hornoch, K. 2012, Icarus, 221, 365
Tedesco, E. F., Noah, P. V., Noah, M., & Price, S. D. 2002, The Astronomical Journal, 123,
1056
Tholen, D. J., Mainzer, A. K., Bauer, J. M., et al. 2014, Minor Planet Electronic Circulars,
145
Walsh, K. J., Delb´o, M., Bottke, W. F., Vokrouhlick´y, D., & Lauretta, D. S. 2013, Icarus,
225, 283
Williams, G. V. 2012, PhD thesis, Open University UK
Wolters, S. D., Green, S. F., McBride, N., & Davies, J. K. 2005, Icarus, 175, 92
—. 2008, Icarus, 193, 535
Wright, E. L. 2007, ArXiv Astrophysics e-prints
Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al. 2010, The Astronomical Journal,
140, 1868
40
Zellner, B. 1979, Asteroid taxonomy and the distribution of the compositional types, ed.
T. Gehrels, 783–806
This manuscript was prepared with the AAS LATEX macros v5.2.
41
E1
E1NOTE TO EDITOR: Complete Tables 1, 4, and 5 should appear in online supplementary
material.
42
|
1110.3935 | 1 | 1110 | 2011-10-18T11:09:37 | Variation of Area-to-Mass-Ratio of HAMR Space Debris Objects | [
"astro-ph.EP"
] | An unexpected space debris population has been detected in 2004 Schildknecht et al. (2003, 2004) with the unique properties of a very high area-to-mass ratio (HAMR) Schildknecht et al. (2005a). Ever since it has been tried to investigate the dynamical properties of those objects further. The orbits of those objects are heavily perturbed by the effect of direct radiation pressure. Unknown attitude motion complicates orbit prediction. The area-to-mass ratio of the objects seems to be not stable over time. Only sparse optical data is available for those objects in drift orbits. The current work uses optical observations of five HAMR objects, observed over several years and investigates the variation of their area-to-mass ratio and orbital parameters. A normalized orbit determination setup has been established and validated with two low and two of the high ratio objects, to ensure, that comparable orbits over longer time spans are determined even with sparse optical data. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 10 (2011)
Printed 31 October 2018
(MN LATEX style file v2.2)
Variation of Area-to-Mass-Ratio of HAMR Space Debris
Objects
C. Fruh1 ⋆† and T. Schildknecht1
1Astronomical Institute, University of Bern, Sidlerstrasse 5, 3012 Bern, Switzerland
Accepted 2011 October 12. Received 2011 October 11; in original form 2011 August 5, the definitve version is available at www.blackwell-
synergy.com
ABSTRACT
An unexpected space debris population has been detected in 2004 Schildknecht et al.
(2003, 2004) with the unique properties of a very high area-to-mass ratio (HAMR)
Schildknecht et al. (2005a). Ever since it has been tried to investigate the dynamical
properties of those objects further. The orbits of those objects are heavily perturbed
by the effect of direct radiation pressure. Unknown attitude motion complicates orbit
prediction. The area-to-mass ratio of the objects seems to be not stable over time.
Only sparse optical data is available for those objects in drift orbits.
The current work uses optical observations of five HAMR objects, observed
over several years and investigates the variation of their area-to-mass ratio and
orbital parameters. A normalized orbit determination setup has been established and
validated with two low and two of the high ratio objects, to ensure, that comparable
orbits over longer time spans are determined even with sparse optical data.
Key words:
analysis
celestial mechanics -- catalogues -- space debris -- observational data
1
INTRODUCTION
The Astronomical
Institute of the University of Bern
(AIUB) detected high area-to-mass ratio (HAMR) objects
in GEO-like orbits in 2004 Schildknecht et al. (2003, 2004,
2005a). Since then, the AIUB observes HAMR objects
on a regular basis and keeps a small catalog of HAMR
and other space debris objects, which are not listed in the
USSTRATCOM catalog. The observations are performed
with the one meter ESA Space Debris Telescope (ESASDT),
located on Tenerife, Spain, and the one meter Zimmerwald
Laser and Astrometry Telescope (ZIMLAT),
located in
Zimmerwald, Switzerland. Additional observations for some
objects, which were detected by the AIUB, are provided by
courtesy of the Keldysh Institute of Applied Mathematics,
Moscow, via the ISON network.
Maintaining a catalogue of HAMR objects is especially
challenging due to the unique properties of these objects;
the orbits are highly perturbed by direct radiation pressure.
Regular observations on short time intervals are mandatory.
In routine orbit determination for catalogue maintenance,
⋆ E-mail:
[email protected] (TS)
† corresponding author
[email protected]
(CF);
variations in the value of the effective area-to-mass ratio
(AMR) were detected, first investigations were performed
Musci et al. (2010).
For
the investigations presented in this paper orbits
are determined with an enhanced version of the CelMech
tool Beutler (2005). The area-to-mass (AMR) value is
determined as a scaling parameter of the direct radiation
pressure. The acceleration due to the direct radiation
pressure is calculated as:
~arad =
C
2
·
S
c
·
AU 2
~r − ~r⊙2 ·
A
m
·
~r − ~r⊙
~r − ~r⊙
,
(1)
where ~r is the geocentric position of the satellite, ~r⊙
the geocentric position vector of the sun, AU the as-
tronomical unit, A the effective cross section exposed
to the radiation, m the mass of the satellite, and c the
speed of light. C is the reflection coefficient. The direct
radiation pressure is determined under the assumption of
a spherically shaped object. In contrast to the calculation
of the radiation pressure acceleration by other sources
(compare e.g. Vallado & McCain (2001)), the coefficient C
is divided by two in the formula above. A value for C has
to be chosen, by default, 2.0 is selected in the standard
processing. This corresponds to an assumption of
full
absorption. All AMR values presented in this paper have
2
C. Fruh and T. Schildknecht
Table 1. Internal name, eccentricity, inclination (deg),
semi-major axis (km), area-to-mass ratio (m2/kg) and
apparent magnitude (mag) of the selected objects of the
AIUB catalogue
observations are allowed per set. An observation set may
consist of more than one tracklet. But the observations
within the sets should not be distributed over more than
three days.
NAME
Epoch
a
e
i
AMR Mag
E03174A 55208.0
E06321D 55275.9
E06327E
54470.1
E08241A 55213.0
41900
41400
40000
41600
0.001
0.035
0.067
0.041
10.1
7.00
12.31
13.26
0.01
2.29
0.20
1.24
14.6
15.3
17.2
16.1
to be interpreted as the effective area-to-mass ratio scaled
with C/2 = 1; the AMR values of other sources may be
scaled with a different factor. It is assumed that the AMR is
constant over the orbital fit interval. A default value of 0.02
m2kg−1 is selected, which corresponds to an AMR value
of a standard GPS satellite, in case the AMR parameter
is not estimated but kept fixed in the orbit determina-
tion. For HAMR objects always an AMR value is estimated.
The shadow paths of
the orbit are modeled, under
the assumption of a spherical earth on a mean circular
orbit; the boundary between sunlit and eclipsed part is
assumed to be cylindrical, no distinction between penumbra
and umbra is made, earth atmosphere is neglected.
For a long term investigation of the orbits and the
AMR values, different comparable orbits have to be deter-
mined. Only sparse observations are available, which are
unequally spaced in time. A normalized setup is developed,
tested with two low AMR objects and two of the HAMR
objects of the AIUB catalog and applied for the creation
of comparable orbits for the investigation of the HAMR
objects.
2 NORMALIZED SPARSE DATA SETUP
2.1 The Method
Four representative GEO objects from the internal cata-
logue of the AIUB were chosen, they have been followed over
longer time periods and are not listed in the USSTRAT-
COM catalogue. Those objects are clearly space debris,
since no maneuvers could be detected in the data. The
AIUB did not have information what those objects actually
were before becoming debris. From the apparent magnitude
it can be concluded that those are all fragmentation pieces.
They represent typical objects found in GEO surveys. Their
properties are listed in Tab. 1.
Orbits were determined for different spacings of
two
observation sets stemming a) from one observation site only
and b) from different sites. In the first case, the observations
either stem from ZIMLAT or from ESASDT only. In the
second case, not only the observations of ZIMLAT and
ESASDT were combined but also observations of the ISON
network, if available. When observations from different sites
are used in orbit determination, the distribution is either
that the first set of observations stems from one site and the
second from another, or that there are observations from
different sites at similar epochs used within the first and/or
the last set of observations or a mixture of those options.
In the figures the label ALL is applied, when observations
of ZIMLAT (labeled ZIM ), the ESASDT and of the ISON
network are combined; the label SDT-ZIM is applied, if
only the observations of ZIMLAT and the ESASDT are
used. The distances between the observations and the
ephemerides of the predicted orbits of the four objects for a
prediction interval of 50 days after the last observation used
for orbit determination were determined. The distances
were averaged and a mean value and standard deviation
was calculated. Between six and 50 single distances between
ephemerides and observations were averaged.
The predicted ephemeris positions are
compared to
the optical angle-only observations, which were not used
in orbit determination. Angular distances are determined
on the celestial sphere. The observation used for the
comparison stem from ZIMLAT and ESASDT and serve as
ground truth. Calibration measurements with high accuracy
ephemerides of Global Navigation Satellite System (GNSS)
satellites provided by International GNSS Service (IGS)
showed an accuracy of the measurements of ZIMLAT
and ESASDT of below one arcsecond. That the further
observations in fact belong to the same object is validated
via an orbit determination with both the observations used
in the original sparse data orbit determination and the
observations, which they were compared to. An orbit deter-
mination with a root-mean-square of below two arcseconds
is a reliable tool to associate observations of this accuracy
of the same object to each other, as shown with cluster
observations in Musci et al. (2005).
2.2 Results
Two of the objects have low area to mass ratios, two objects
qualify as HAMR objects with an AMR value larger than
1m2/kg. The optical angle-only observations are obtained
with ZIMLAT (Zimmerwald, Switzerland), and ESASDT
(Tenerife, Spain), supplemented by some observations of the
ISON network provided by the Keldysh Institute of Applied
Mathematics, Moscow, Russia. The latter observations were
obtained from different sites of the ISON network, in these
particular cases, all located in Eastern Europe.
In Fig. 1 the angular distance between predicted and
observed position are displayed as a function of the time
interval between the first and the last observation, which
were used in orbit determination. Displayed are the mean
values and the standard deviations of the angular distances
of the single orbits. The mean value and standard devia-
tions are determined with the single angular distances of
predicted position to observed ones, all within 50 days since
orbit determination.
All orbits were determined from two observation sets
only, using a priori orbital elements. A maximum of eight
Figure 1 shows that the angular distances are in gen-
eral very small. The vast majority of the determined orbits
Variation of Area-to-Mass-Ratio of HAMR Space Debris Objects
3
[ht]
(a)
(b)
(c)
(d)
Figure 1. Angular distances as a function of the time interval between the first and the last observation of the fit
interval of orbit determination for object (a) E03174A, (b) E06321D, (c) E06327E and (d) E08241A.
[h]
(a)
(b)
(c)
(d)
Figure 2. Root mean square of orbit determination as a function of the arc length of observations for object (a)
E03174A, (b) E06321D, (c) E06327E and (d) E08241A.
[h]
(a)
(b)
(c)
(d)
Figure 3. Angular distances as a function of the number of observations used for orbit determination for object (a)
E03174A, (b) E06321D, (c) E06327E and (d) E08241A.
[h]
(a)
(b)
(c)
(d)
Figure 4. Angular distances as a function of anomaly distribution factor for object (a) E03174A, (b) E06321D, (c)
E06327E and (d) E08241A.
4
C. Fruh and T. Schildknecht
[h]
(a)
(b)
(c)
(d)
Figure 5. Angular distance as a function of time interval covered by observations used for orbit determination for
object (a) E03174A, (b) E06321D, (c) E06327E and (d) E08241A.
[h]
(a)
(b)
(c)
(d)
Figure 6. Time interval covered by the observations within the sets as a function of the number of observations used
for orbit determination for object (a) E03174A, (b) E06321D, (c) E06327E and (d) E08241A.
even produce distances smaller than 0.6 degrees. Except
for the first object, each object also shows some outliers,
with larger angular distances. These larger distances also
tend to show larger standard deviations. The value of the
angular distances seems to be, at least in this setup, quite
independent of how large the difference between the first
and the last observations of the fit interval is. Moreover,
Figure 1 also shows that there is no significant difference in
using observations only from one observation site for orbit
determination or using observations from different sites.
It could not be shown that the latter approach is more
advantageous for orbit determination. Different observation
sites still have advantages in terms of availability, weather
conditions, which results in a larger amount of observations,
which are available. In Fig. 2, the root mean square of the
orbit determinations is shown, which were used for the
prediction, as a function of the angular distance. No trend
is visible, all orbits, which were determined had a small
root mean square of below three arcseconds.
In Fig. 3,
the angular distances are displayed as a
function of the actual number of single observations that
entered orbit determination. It can be seen that no strong
correlation is visible between the actual number of observa-
tions used and the value for the distances.
To find a measure for the true anomaly distribution,
an anomaly distribution measure fano was defined: It would
be ideal to distribute all n observations equally spaced with
an angle of 2π/n between each observation. The deviation
from this ideal distribution is determined and normalized
with the number of observations. The smaller fano, the
better distributed are the observations in anomaly.
fano =
1
n
vuut
n−1
X
i=1 (cid:0)
2π
n
− (ai+1 − ai)(cid:1)2
+ (cid:0)
2π
n
− (a1 + 2π − an)(cid:1)2
,
(2)
where as n is the number of observations and ai with
i = 1, .., n are the anomalies of the single observations,
in ascending anomaly order. The angular distances as
a function of fano are displayed in Fig. 4. There is no
clear correlation between the fano and the distances, as
it is expected for objects with small eccentricities. Object
E06327E, with the highest eccentricity of e=0.06, has the
strongest correlation with fano.
The crucial
factor however, seems to be the time in-
terval covered by observations itself within the sets. In
Fig 5, the angular distances are displayed as a function of
the time interval covered within the two sets used in the
beginning and the end of the fit interval, without the time
gap in between the two sets. A strong correlation is visible.
Fig. 6 shows that there is no strong correlation between
the number of used observations and the time interval
covered within the sets. For example for the ESASDT
observation strategy, primarily densely spaced observations
are available.
The investigation of the data displayed in Fig. 6 showed
that a coverage of at least 1.2 hours for both sets together
seems to be necessary, in order to gain an orbit which allows
to safely re-detect the investigated objects in more than
Variation of Area-to-Mass-Ratio of HAMR Space Debris Objects
5
[ht]
(a)
(b)
(c)
Figure 7. Inclination as a function of time for orbits of the object (a)E08241A, (b) E06321D, (c) E07194A, (d) E07308B, (e) E06293A.
(d)
(e)
[h]
(a)
(b)
(c)
Figure 8. Eccentricity as a function of time for orbits of the object (a) E08241A, (b) E06321D, (c) E07194A, (d) E07308B, (e) E06293A.
(d)
(e)
90 percent of all cases with a field of view of one square
degree, that is to have an accuracy of below 0.5 degrees.
3
INVESTIGATION OF HAMR OBJECTS IN
SPARSE DATA SETUP
The dynamical properties of HAMR objects were studied in
the normalized sparse data setup established in the previous
section. Orbits are determined with two observation sets
only. The sets consist of four to eight observations each. The
observations are required to span at least a time interval
of 1.2 hours within the sets and need to be well spread
over the anomaly for the objects in orbits with a high
eccentricity. The total fit interval for orbit determination
ranges between 10 and 120 days. As shown in the previous
section the comparability of the orbits do not seem to be
dependent on these ranges.
The orbits were first determined with observations from one
observation site only, then with observations from different
sites in the setup mentioned above. The observations used
in this investigation stem from the ESASDT, ZIMLAT, and
from several telescopes of the ISON network.
6
C. Fruh and T. Schildknecht
[ht]
(a)
(b)
(c)
Figure 9. AMR as a function of time for orbits of the object (a) E08241A, (b) E06321D, (c) E07194A, (d) E07308B, (e) E06293A.
(d)
(e)
[h]
(a)
(b)
Figure 10. Relative variation of AMR value as a function of the absolute AMR value of (a) 47 HAMR and (b) LAMR objects.
3.1 Selected Objects
Five objects were selected for a detailed investigation. All
objects were discovered and first detected by the AIUB and
are not listed in the USSTRATCOM catalogue. All objects
are faint debris objects. They were tracked successfully over
several years, and no maneuvers were detected. A set of
osculating orbital elements and an average value for the
apparent magnitudes are listed in Tab. 2. The two objects
with the lowest AMR values, E08241A and E06321D, which
were used in the investigation of the sparse data orbit
determination, are used here again.
3.2 Evolution of Orbital Elements
The evolution of the orbital elements over time is in-
spected in a first step. Figure 7 shows the development
of the inclination and its errors in inclination, of the
five objects. The error bars are too small, to be visible
in the plot in most cases. The inclination values of the
Table 2. Investigated HAMR objects: Internal name,
epoch (MJD), eccentricity, inclination (deg), semi-major
axis (km), area to mass ratio (m2/kg) and apparent mag-
nitude (mag).
NAME
Epoch
a
e
i
AMR Mag
E08241A 55213.0
E06321D 55275.9
E07194A 54877.0
E07308B
54416.0
E06293A 54951.0
41600
41400
40900
35600
40200
0.041
0.035
0.005
0.264
0.245
13.26
7.00
7.31
7.63
11.06
1.24
2.29
3.37
8.83
15.41
16.1
15.3
16.8
15.8
16.8
different orbits are closely aligned to each other and mark
a consistent evolution, only in the case of object E08241A
in Fig. 7 a wider spread in the inclination values can be
observed. The orbits determined with observations from
the different observation sites produce almost identical
results. For object E07308B and E06293A, which have the
highest AMR values, the inclination seems not to follow
a steady increase over time, but some smaller periodic
substructure seems to be superimposed. These may very
well be the perturbations with a period of one nodal
Variation of Area-to-Mass-Ratio of HAMR Space Debris Objects
7
[h]
(a)
(b)
(c)
Figure 11. Angular distances of predicted orbits on the celestial sphere as a function of AMR for orbits of the object (a) E08241A, (b)
E06321D, (c) E07194A, (d) E07308B, (e) E06293A.
(d)
(e)
[h]
(a)
(b)
(c)
Figure 12. Error of the AMR value as a function of AMR as estimated in orbits of the object (a) E08241A, (b) E06321D, (c) E07194A,
(d) E07308B, (e) E06293A.
(d)
(e)
year, which are well known for objects with high AMR,
see e.g., Liou & Weaver (2005), Schildknecht et al. (2005b).
3.3 Evolution of Area to Mass Ratio Value
Figure 8 shows the evolution of the eccentricity values
and its errors estimated in orbit determination for the
different objects. Periodic variations can be observed for all
objects. The different orbits with observations from one site
only or from different sites result in the same eccentricities.
Figure 9 shows the AMR values as a function of time for
the objects listed in Tab. 2. In all cases, the values for the
AMR do not show clear and obvious common trends, see
Fig. 7 and 8.
For object E08241A, the AMR values vary around a
mean value of 1.4 m2kg−1 with no obvious trend or periodic
signal, see Fig. 9a.
For object E06321D (see Fig. 9b), the AMR value seems to
vary periodically with a period of about one year around
8
C. Fruh and T. Schildknecht
[h]
(a)
(b)
(c)
Figure 13. Absolute values and standard deviations of the angular distances as a function of the error of the AMR value as found in
orbit determination of the object (a) E08241A, (b) E06321D, (c) E07194A, (d) E07308B, (e) E06293A.
(d)
(e)
a value of 2.5 m2kg−1, but also values of 2.35 m2kg−1
and 2.65 m2kg−1 occur. Similar results were obtained by
Musci et al. (2010), for the same object, in different orbit
determination setups. The AMR value of object E07194A
(see Fig. 9c) varies around 3.5 m2kg−1, but in the orbits
determined with combined observations from all the sites,
so-called outliers of 4.5 m2kg−1 and 2.3 m2kg−1 occur as
well. These have, however large error values.
to generally in-
Object E07308B (see Fig. 9d)
crease its AMR value over time from a value of 8.5 m2kg−1
up to 9.0 m2kg−1. But single orbits also show AMR values
of i.e. 10 m2kg−1.
seems
that object E06293A, which is
the
Figure 9e shows
object with the largest AMR value regarded here, has
significant data gaps. A general trend of the AMR value in
time, increasing from 15.5 m2kg−1 to 16.5 m2kg−1 cannot
be excluded. But one orbit determined with ESASDT data
also shows a value of 18.2 m2kg−1, with a small formal error.
No general correlation between the AMR value itself
and the variations of the AMR value could be determined,
no general trend is visible. A study on the variation
of AMR values was conducted by Schildkecht et al.
(to be published 2011). The variations of the AMR values
of 47 HAMR objects were investigated and compared to
the AMR variations of orbits of 40 low AMR (LAMR)
value objects. No normalized or sparse data setup orbit
determination setup was chosen. The AMR values in that
analysis were determined in the standard orbit determina-
tion procedure at the AIUB, with fit arcs as long as possible
for a successful, that is defined as leading to a small rms
error, orbit determination. The results are illustrated in
Fig. 10. No general trend in the AMR variations could
be determined for either HAMR or LAMR objects. The
relative variations of the AMR values of the LAMR objects
were larger, than the AMR variations of the HAMR values.
The AMR variations of the LAMR objects were of the
order of several 100 percent.
All orbits were predicted and compared to additional
observations of the same object, which were not used for
orbit determination. The additional observations were all
checked via orbit determination, to ensure that they belong
to the same object. Figure 11 shows the angular distances
between the predicted ephemeris and observations. The
values are averaged over all distances 50 days after orbit
determination and their standard deviations serve as error
bars.
Figure 11a shows that for object E08241A, one orbit
produces the largest distances of one degree. This orbit
does not show up prominently in the orbital parameter
plots (see Fig. 8a and 7a) or AMR value plots (see Fig. 9a).
The orbit with ZIMLAT data, which produced the outlier
AMR value of 0.82m2kg−1, does not show up prominently
in the distance plot (Fig. 11a).
The mean value of all angular distances of object E06321D
are well below 0.2 degrees, but four orbits show large
standard deviations in the angular distance, as Fig. 11b
shows. All of them have been determined with combined
observations from ZIMLAT, ESASDT, and ISON obser-
vations. Their AMR values are 2.36 m2kg−1, 2.50 m2kg−1,
2.57 m2kg−1, and 2.66 m2kg−1. The orbits with the AMR
value of 2.36 m2kg−1 does show up also in a group of
outlier AMR values, which do not seem to follow the
periodic variation in the evolution of the AMR values. The
other orbits, with large standard variations in the angular
distance do not show up prominently (Fig. 9b). Those orbits
with the largest standard variation in angular distance do
not show the largest error in the AMR values either, as
Fig. 13 shows.
Variation of Area-to-Mass-Ratio of HAMR Space Debris Objects
9
shows
Figure 11c
for object E07194A three angular
distances with large standard deviations. The orbits were
determined with observations from all sites. They have
AMR values of 2.12 m2kg−1, 2.21 m2kg−1, and 4.46 m2kg−1.
Those are the smallest and largest AMR values in the deter-
mined orbits for E07194A. These three values do also show
up as outliers in Fig. 9c. For objects E07308B and E06293A,
the angular distances with a large standard variation (see
Fig. 11d and e), do not show significant outlier AMR values
in Fig. 9d and e. For object E07308B, the orbit with an
AMR value of 10.15 m2kg−1 shows the largest mean value
in the angular distance of almost 0.7 degrees but has a small
standard deviation in this distance (Fig. 11d). This value
is significantly different compared to the other determined
AMR values, see Fig. 9d.
The dependency of the AMR value on the error of
the AMR, as it was found in orbit determination,
is
investigated in the final step. No clear correlation could
be determined between the AMR value and its rms value
(Fig. 12).
Figure 13 shows the angular distance distances on the
celestial sphere as a function of the error of the AMR value.
As expected, for none of the objects a clear correlation
between the error of the AMR value and the absolute value
of the distances or the standard deviation of the distances
could be determined.
All
investigated objects show variations in the AMR
value, but not a common characteristic in these variations.
It has to be noted that the result may be affected by the
relatively simple shadowing model that was used in orbit
determination; as it was shown Pardini & Anselmo (2008),
Valk & Lemaıtre (2008), shadowing effects have a signifi-
cant influence on the long term evolution of orbits of HAMR
objects. Investigation of simulated orbits with numerical
and semi-analytical methods, e.g. by Valk & Lemaıtre
(2008), Valk et al.
(2008) also showed the existence of
irregular chaotic orbits and the significant influence of
secondary resonances on the orbits of HAMR objects.
However, these simulations did assume a constant AMR
value. Complex attitude motion, irregular shapes, and/or
deformation of the actual objects, could lead to an actual
change in the AMR value itself over time, which may not
be averaged out over the fit interval of orbit determination.
4 CONCLUSIONS
A sparse data setup was established to create compa-
rable orbits over longer time intervals. Orbits with two
data sets only do produce small differences between the
propagated ephemerides and further observations, as long
as 1.2 hours are covered within the sets. Other factors,
such as that the observations stem from different sites or
the time interval between the sets, are found to be negligible.
The orbits of high area-to-mass ratio (HAMR) objects
were analyzed in this setup. The AMR value, that is
the scaling factor of the direct radiation pressure (DRP)
parameter, varies over time. The order of magnitude of the
variation of the area-to-mass ratio (AMR) value was not
correlated with the order of magnitude of its error.
The variation of the AMR is not averaged out in the
fit interval of orbit determination. In the evolution of the
AMR value over time, no common characteristic could
be determined for different HAMR objects. Further work
on the orbits of HAMR objects is needed, to improve the
radiation pressure model, to determine possible attitude
motion or deformations and to understand also resonance
effects and the existence of chaotic regions.
ACKNOWLEDGMENTS
Special thank goes to ISON and the Keldysh Institute
of Applied Mathematics, Moscow, for the supplementary
observations.
The work was supported by the Swiss National Science
Foundation through grants 200020-109527 and 200020-
122070.
The observations
under ESA/ESOC contracts
17835/03/D/HK.
The authors thank the reviewer for useful hints.
from the ESASDT were acquired
15836/01/D/HK and
REFERENCES
Beutler, G. 2005. Methods of Celestial Mechanics. Two Vol-
umes. Springer-Verlag, Heidelberg. ISBN: 3-540-40749-9
and 3-540-40750-1.
Liou, J.-C., & Weaver, J.K. 2005. Orbital Dynamics of High
Area-to-Mass Ratio Debris and Their Distribution in the
Geosynchronos Region. In: Proceedings of the Forth Eu-
ropean Conference on Space Debris, pp. 119-124, ESOC,
Darmstadt, Germany, 18-20 April 2005.
Musci, R., Schildknecht, T. Flohrer, T. & Beutler, G. 2005.
Concept for a Catalogue of Space Debris in GEO.
In:
Proceedings of the Fourth European Conference on Space
Debris, pp. 601-606, ESOC, Darmstadt, Germany, 18-20
April 2005.
Musci, R., Schildknecht, T. & Ploner, M. 2010. Analyzing
long Observation Arcs for Objects with high Area-to-Mass
Ratios in Geostationary Orbits.
In: Acta Astronautica,
vol. 66, pp 693-703.
Pardini, C., & Anselmo, L. 2008. Long-Term Evolution of
Geosynchronous Orbital Debris with High Area-to-Mass
Ratios. Trans. Japan Soc. Aero. Space Sci., 51, 22 -- 27.
Schildkecht, T., Fruh, C. Hinze, A. & Herzog, J.to be pub-
lished 2011. Dynamical Properties of High Area to Mass
Ratio Objects in GEO-Like Orbits. Advances in Space
Research.
Schildknecht, T., Musci, R. Ploner, M. Flury, W. Kuusela,
J. de Le´on Cruz, J. & de Fatima Dom´ınguez Palmero,
L. 2003. An Optical Search for Small-Size Debris in GEO
and GTO. In: Proceedings of the 2003 AMOS Technical
Conference, 9-12 September 2003, Maui, Hawaii, USA.
Schildknecht, T., Musci, R. Ploner, M. Beutler,
G. Kuusela, J. de Le´on Cruz, J. & de Fatima
Dom´ınguez Palmero, L. 2004.
Optical Observations
10
C. Fruh and T. Schildknecht
of Space Debris in GEO and in Highly-Eccentric Orbits.
Advances in Space Research, 34(5), 901 -- 911.
Schildknecht, T., Musci, R. Flury, W. Kuusela, J.
de Le´on Cruz, J. & de Fatima Dom´ınguez Palmero,
L. 2005a. Optical Observations of Space Debris in High-
Altitude Orbits.
In: Proceedings of the Forth European
Conference on Space Debris, pp. 113-118, ESOC, Darm-
stadt, Germany, 18-20 April 2005.
Schildknecht, T., Musci, R. Flury, W. Kuusela, J.
de Le´on Cruz, J. & de Fatima Dom´ınguez Palmero,
L. 2005b. Properties of the High Area-to-Mass Ratio Space
Debris Population in GEO. In: Proceedings of the 2005
AMOS Technical Conference, 5-9 September 2005, Maui,
Hawaii, USA.
Valk, S., & Lemaıtre, A. 2008. Semi-Analytical Investiga-
tions of High Area-to-Mass Ratio geosynchronous Space
Debris Including Earth's Shadowing Effects. Advances in
Space Research, 42(8), 1429 -- 1443.
Valk, S., Delsate, N. Lemaıtre, A. & Carletti, T. 2008.
Semi-Analytical Investigations of High Area-to-Mass Ra-
tio geosynchronous Space Debris Including Earth's Shad-
owing Effects. Advances in Space Research, 43(10), 1509 --
1526.
Vallado, D., & McCain, W. 2001. Fundmentals of Astrody-
namics and Applications. Microcosm Press, El Segundo,
California. ISBN 0-7923-6903-3.
|
1211.6365 | 1 | 1211 | 2012-11-27T17:25:46 | AKARI/IRC 18 Micron Survey of Warm Debris Disks | [
"astro-ph.EP",
"astro-ph.SR"
] | Context. Little is known about the properties of the warm (Tdust >~ 150 K) debris disk material located close to the central star, which has a more direct link to the formation of terrestrial planets than the low temperature debris dust that has been detected to date. Aims. To discover new warm debris disk candidates that show large 18 micron excess and estimate the fraction of stars with excess based on the AKARI/IRC Mid-Infrared All-Sky Survey data. Methods. We have searched for point sources detected in the AKARI/IRC All-Sky Survey, which show a positional match with A-M dwarf stars in the Tycho-2 Spectral Type Catalogue and exhibit excess emission at 18 micron compared to that expected from the Ks magnitude in the 2MASS catalogue. Results. We find 24 warm debris candidates including 8 new candidates among A-K stars. The apparent debris disk frequency is estimated to be 2.8 +/- 0.6%. We also find that A stars and solar-type FGK stars have different characteristics of the inner component of the identified debris disk candidates --- while debris disks around A stars are cooler and consistent with steady-state evolutionary model of debris disks, those around FGK stars tend to be warmer and cannot be explained by the steady-state model. | astro-ph.EP | astro-ph | Astronomy&Astrophysicsmanuscript no. ms
November 1, 2018
c(cid:13) ESO 2018
AKARI/IRC 18 µm Survey of Warm Debris Disks
Hideaki Fujiwara1, Daisuke Ishihara2, Takashi Onaka3, Satoshi Takita4, Hirokazu Kataza4, Takuya Yamashita5,
Misato Fukagawa6, Takafumi Ootsubo7, Takanori Hirao8, Keigo Enya4, Jonathan P. Marshall9, Glenn J. White10,11,
Takao Nakagawa4, and Hiroshi Murakami4
2
1
0
2
v
o
N
7
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
6
3
6
.
1
1
2
1
:
v
i
X
r
a
1 Subaru Telescope, National Astronomical Observatory of Japan, 650 North A'ohoku Place, Hilo, HI 96720, USA
e-mail: [email protected]
2 Graduate School of Science, Nagoya University, Furo-cho, Chikusa-ku, Nagoya, Aichi 464-8602, Japan
3 Department of Astronomy, School of Science, University of Tokyo, Bunkyo-ku, Tokyo 113-0033, Japan
4 Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1 Yoshinodai, Chuo-ku, Sagamihara,
Kanagawa 252-5210, Japan
5 National Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-0015, Japan
6 Graduate School of Science, Osaka University, 1-1 Machikaneyama, Toyonaka, Osaka 560-0043, Japan
7 Astronomical Institute, Tohoku University, 6-3 Aramaki, Aoba-ku, Sendai, Miyagi 980-8578, Japan
8 Research Institute of Science and Technology for Society, Japan Science and Technology Agency, K's Gobancho Bldg, 7,
Gobancho, Chiyoda-ku, Tokyo 102-0076, JAPAN
9 Departmento F´ısica Te´orica, Facultad de Ciencias, Universidad Aut´onoma de Madrid, Cantoblanco, 28049 Madrid, Spain
10 Department of Physics and Astronomy, The Open University, Walton Hall, Milton Keynes, MK7 6AA, UK
11 Space Science & Technology Department, The Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire OX11 0QX, UK
Received 19 June 2012 / Accepted 21 November 2012
ABSTRACT
Context. Little is known about the properties of the warm (Tdust & 150 K) debris disk material located close to the central star, which
has a more direct link to the formation of terrestrial planets than the low temperature debris dust that has been detected to date.
Aims.To discover new warm debris disk candidates that show large 18 µm excess and estimate the fraction of stars with excess based
on the AKARI/IRC Mid-Infrared All-Sky Survey data.
Methods. We have searched for point sources detected in the AKARI/IRC All-Sky Survey, which show a positional match with A-M
dwarf stars in the Tycho-2 Spectral Type Catalogue and exhibit excess emission at 18 µm compared to that expected from the KS
magnitude in the 2MASS catalogue.
Results. We find 24 warm debris candidates including 8 new candidates among A-K stars. The apparent debris disk frequency is
estimated to be 2.8 ± 0.6%. We also find that A stars and solar-type FGK stars have different characteristics of the inner component
of the identified debris disk candidates -- while debris disks around A stars are cooler and consistent with steady-state evolutionary
model of debris disks, those around FGK stars tend to be warmer and cannot be explained by the steady-state model.
Key words. circumstellar matter -- zodiacal dust -- infrared: stars
1. Introduction
Some main-sequence stars are known to have dust disks around
them. Primordial dust grains that originally exist in protoplane-
tary disks around T Tauri stars or Herbig Ae/Be stars are dissi-
pated before their central stars reach the main-sequence phase.
Therefore the circumstellar dust grains around main-sequence
stars should be composed of second generation dust grains re-
plenished during the main-sequence phase, rather than primor-
dial dust from protoplanetary disks. These second generation
dust grains are thought to have originated from collisions of
planetesimals, or during the destruction of cometary objects (e.g.
Backman & Paresce 1993; Lecavelier Des Etangs et al. 1996),
giving a reason why circumstellar dust disks around main-
sequence stars are named "debris disks." Debris disks are ex-
pected to be related to the stability of minor bodies, and poten-
tially the presence of planets around stars (e.g. Wyatt 2008).
Debris disks are identified from the spectral energy distri-
butions (SEDs) of stars that show an excess over their expected
photospheric emission at infrared (IR) wavelengths, since cir-
cumstellar dust grains absorb the stellar light and re-emit mainly
in the IR region. The first example of the debris disk, around
Vega, was discovered through its excess emission at wavelengths
of & 60 µm by Infrared Astronomical Satellite (IRAS) in 1980s
(Aumann et al. 1984). After the discovery of the Vega debris
disk, more than 100 other debris disks were identified from
the IRAS catalogue. A number of new debris disk candidates
have been discovered recently, even though the IRAS mission
flew more than 20 years ago (Rhee et al. 2007). The power of
a next generation all-sky survey satellite mission like AKARI
(Murakami et al. 2007) to take forward the debris disk search
is evident.
Most of the known debris disks only show excess far-infrared
(FIR) emission at wavelengths longer than 25 µm. This excess
comes from the thermal emission of dust grains with low tem-
peratures (Tdust ∼ 100 K), and is an analogue of Kuiper belt
objects in the Solar System. Recently Herschel (Pilbratt et al.
2010) has provided FIR data for debris disks with high sensitiv-
ity and high spatial resolution and is opening a new horizon of
research on cold debris disks (e.g. Thompson et al. 2010). For
example, Eiroa et al. (2011) discovered cold (Tdust . 20 K),
faint debris disks towards three GK stars, which might be in a
1
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
new class of coldest and faintest disks discovered so far around
mature stars and which cannot easily be explained by invoking
classical debris disk models. In addition, some cold debris disks
have been spatially resolved by Herschel (e.g. Lohne et al. 2012;
Liseau et al. 2010). On the other hand, little is known to date
about the warm (Tdust & 150 K) debris disk material located
close to the star, which should be an analogue of the asteroid
belt in the Solar System. Warm dust grains in the inner region
of debris disks should have a more direct link to the formation
of terrestrial planets than the low temperature dust that has been
previously studied (Meyer et al. 2008). Recent high-sensitivity
surveys at 10 − 20 µm allow us to investigate the properties of
this inner debris disk material.
AKARI is a Japanese IR satellite dedicated primarily to an
IR all-sky survey (Murakami et al. 2007). It was launched in
February 2006. The Mid-Infrared (MIR) All-Sky Survey was
performed using 9 and 18 µm broad band filters with the
InfraRed Camera (IRC; Onaka et al. 2007) onboard AKARI un-
til August 2008 (Ishihara et al. 2010), and a complementary
FIR survey at 65 -- 160 µm was performed with the Far-Infrared
Surveyor (Yamamura et al. 2010). We show that the data from
the AKARI/IRC All-Sky Survey is very powerful to identify
warm debris disks and to contribute to increase the number of
sample with its higher sensitivity and spatial resolution than
those of IRAS.
Here we report initial results of survey of warm debris disks
around main-sequence stars based on the AKARI/IRC All-Sky
Survey.
2. 18 µm Excess Search with AKARI/IRC All-Sky
Survey
2.1. InputCatalogue:Tycho-2SpectralTypeCatalogue
In this work, we use the Tycho-2 Spectral Type Catalogue
(Wright et al. 2003) as an input catalogue. The Tycho-2 Spectral
Type Catalogue lists spectral types for 351864 of the Tycho-
2 stars by cross-referencing Tycho-2 to several catalogues
that have spectral
types (the Michigan Catalogue for the
HD Stars, Volumes 1-5 (Houk & Cowley 1975; Houk 1978,
1982; Houk & Smith-Moore 1988; Houk & Swift 1999); the
Catalogue of Stellar Spectra Classified in the Morgan-Keenan
System (Jaschek et al. 1964); the MK Classification Extension
(Kennedy 1996); the Fifth Fundamental Catalogue (FK5) Part
I (Fricke et al. 1988) and Part II (Fricke et al. 1991); and the
Catalogue of Positions and Proper Motions (PPM), North
(Roeser & Bastian 1988) and South (Bastian & Roser 1993)),
using the VizieR astronomical database1.
We selected stars whose luminosity classes were labeled as
"V" (dwarf) from the catalogue to remove giant stars from the
sample, since we focus our search for debris disks around main-
sequence stars. Since young and massive main-sequence stars,
which are commonly known as classical Be stars, sometimes
exhibit significant IR excess resulting from free-free emission
from circumstellar gas disks, stars whose spectral types were la-
belled as "O" or "B" in the Tycho-2 Spectral Type Catalogue
were excluded from our sample. The selection of A-M dwarfs
from the Tycho-2 Spectral Type Catalogue left 64209 stars as
an input catalogue for further investigation. Since the V-band
limiting magnitude in the Tycho-2 Catalogue is about V = 11.5
(Hog et al. 2000), the distance limit for A0, F0, G0, K0, and M0
dwarfs probed by the Tycho-2 Catalogue is estimated as 1500,
1 URL: http://vizier.u-strasbg.fr/viz-bin/VizieR/.
2
580, 260, 130, and 35 pc, respectively, by assuming no extinction
at the V-band.
The spatial distributions of the A-M dwarf stars from the
Tycho-2 Spectral Type Catalogue in equatorial coordinates are
shown in Figure 1. Since the information on luminosity classes
in the Tycho-2 Spectral Type Catalogue has been adopted mainly
from the Michigan Catalogue, which was based on the observa-
tions taken at Cerro Tololo Inter-American Observatory with the
Michigan Curtis Schmidt telescope, the spatial distribution of
the A-M dwarf stars used in this study is unevenly distributed in
the southern hemisphere.
2.2. AKARI/IRCPointSourceCatalogue
In the AKARI All-Sky Survey, the IRC was operated in the scan
mode with a scan speed of 215′′ s−1 and the data sampling time
of 0.044 s, which provided a spatial sampling of ∼ 9.′′4 along the
scan direction. The pixel scale along the cross-scan direction was
∼ 9.′′4 by binning 4 pixels in a row to reduce the data downlink
rate (Ishihara et al. 2010). The positional accuracy is improved
to ∼ 5′′ in both bands by combining the dithered data from multi-
ple observations using the standard AKARI/IRC All-Sky Survey
pipeline software (Ishihara et al. 2010). The 5 σ sensitivity for
a point source per scan is estimated to be 50 mJy in the S9W
band and 90 mJy in the L18W band. The distance limit at which
the stellar photosphere can be detected in the L18W band of the
AKARI/IRC Point Source Catalogue (PSC) is estimated as 74,
40, 26, 17, and 10 pc for A0, F0, G0, K0, and M0 dwarfs, re-
spectively.
The first version of the AKARI/IRC PSC contains more than
850000 sources detected at 9 µm and 195000 sources at 18 µm.
The quoted flux density and its uncertainty for a single source
were calculated as average and deviation (root-mean-square) of
photometry of multiple detections. Details of the AKARI/IRC
All-Sky Survey and its data reduction process are described in
Ishihara et al. (2010).
2.3. Cross-correlation
The selected dwarfs from the Tycho-2 Spectral Type Catalogue
were cross-correlated against the AKARI/IRC PSC positions. A
search radius of 5′′ was adopted to take account of the posi-
tional uncertainty in the AKARI/IRC PSC. After the results of the
cross-correlation, 873 stars out of the input 64209 stars were de-
tected at 18 µm. For the detected stars, we searched for the near-
est near-infrared (NIR) counterparts within a search radius of 5′′
in the 2MASS All-Sky Catalogue of Point Sources (Cutri et al.
2003) and found 856 stars with 2MASS counterparts. The num-
bers of stars that were detected with both of AKARI/IRC 18 µm
and 2MASS are summarised in Table 1.
2.4. 18 µm ExcessCandidateSelection
We made a histogram of the KS−[18] colours of the stars de-
tected at 18 µm for each spectral type and set the KS−[18]
colours as an indicator of 18 µm excess. The histograms of
the KS−[18] colour are shown in Figure 2 for each spectral
type. Gaussian fitting was performed for the distribution of
the KS−[18] colour to determine the thresholds for 18 µm ex-
cess detection. The peak position µKS −[L18W] and profile width
σKS −[L18W] of the KS−[18] colour in each of the spectral types
derived by Gaussian fitting are listed in Table 2. The peak posi-
tion (µKS −[L18W]) corresponds to the empirical KS−[18] colours
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Table 1. Summary of debris disk survey at 18 µm.
Spectral
Type
Input
Sources
18 µm-Detected
Sources
Excess
Sources
Debris
Candidates
A
F
G
K
M
Total
18232
29766
14013
2122
76
64209
196
324
173
144
19
856
21
12
3
2
4
42
11
10
2
1
0
24
Apparent
Frequency
(%)
5.6 ± 1.6
3.1 ± 1.0
1.2 ± 0.8
0.7 ± 0.7
0.0
2.8 ± 0.6
18 µm Sources Debris Candidates Debris Disk
w/ Detectable
Frequency
Photoshpere
w/ Detectable
Photoshpere
(%)
178
311
169
144
19
830
4
3
2
0
0
9
2.1 ± 1.1
1.0 ± 0.6
1.2 ± 0.8
0.0
0.0
1.1 ± 0.4
of normal stars. The scatter in the KS−[18] colours (σKS−[L18W])
are between 0.14 and 0.19 mag. This scatter can be accounted
for by the photometric uncertainties in AKARI/IRC and 2MASS.
For faint stars with KS & 4 mag, the scatter is dominated by the
uncertainty in the AKARI/IRC L18W band photometry, which is
mainly by the background noise (< 0.2 mag), while the uncer-
tainty in the 2MASS KS-band photometry is smaller than 0.02
mag and does not contribute significantly. For bright stars with
KS . 4 mag, the uncertainty in the AKARI/IRC L18W band
photometry is small (. 0.05 mag) and the scatter is dominated
by the uncertainty in the 2MASS KS-band photometry, which
is ∼ 0.1 − 0.2 mag, since those stars are in the range of partial
saturation in the 2MASS observations (Cutri et al. 2003).
We set a 2σ-threshold for a positive 18 µm excess de-
tection for each spectral type as defined in Table 2. Sources
whose KS−[18] colour is larger than the 2σ-thresholds are se-
lected as 18 µm excess candidates. In the selection of the
18 µm excess candidates, the photometric uncertainties of the
individual stars are also taken into account. A visual inspec-
tion of the excess candidates was conducted by comparing
AKARI/IRC images with 2MASS and DSS images to con-
firm that there were no nearby IR sources. During visual in-
spection, HD 68256 and HD 68257, whose separation is 6′′,
HD 24071 and HD 24072, whose separation is 4′′, HD 94602
and HD 94601, whose separation is 7′′, and HD 72945 and
HD 72946, whose separation is 10′′, and HD 147722 and
HD 147723, whose separation is 2′′, were found to be unre-
solved in the AKARI/IRC All-Sky image and could be contam-
inated by each other. Thus they were excluded from considera-
tion as 18 µm excess candidates. BD+213825 is also found to be
contaminated by 2MASS J19334644+2130586, whose separa-
tion is 7′′, and thus excluded. It should be noted that the contam-
inating star 2MASS J19334644+2130586 has a significant MIR
excess, which was confirmed by our follow-up observations us-
ing Subaru/COMICS and Spitzer/IRS. The rejected sources due
to the nearby star contamination are listed in Table 3. Although
HD 89125 and HD 15407 might be contaminated by the nearby
star GJ 387 (separation is 7′′) and HIP 11962 (separation is 20′′),
respectively, the observed excess flux densities at 18 µm are
much larger that the expected flux densities of the contaminating
sources. We conclude that the 18 µm excesses of HD 89125 and
HD 15407 are secure, and we select a final catalogue of 42 stars
with 18 µm excess in the Tycho-2 Spectral Type Catalogue.
The 9 µm flux densities of HD 166191, HD 39415, and
HD 145263, which were selected from the KS−[18] colours,
are not available in the latest version of AKARI/IRC PSC since
they were observed by AKARI at 9 µm during a South Atlantic
Anomaly (SAA) passage, and thus rejected from the PSC.
However the reliable detections are confirmed in the 9 µm im-
ages. We measured the 9 µm flux densities in the images man-
ually. Mouri et al. (2011) investigated the behavior of signals
Table 2. Obtained Parameters from Gaussian Fit.
SpT
A
F
G
K
M
µKS −[L18W]
(mag)
0.187
0.160
0.158
0.212
0.341
σKS −[L18W]
(mag)
0.142
0.174
0.192
0.149
0.179
Threshold(2σ)a
(mag)
0.470
0.545
0.542
0.511
0.700
a Threshold(2σ) = µKS −[L18W] + 2σKS −[L18W].
Table 3. List of Rejected Sources.
Name
HD 24071
HD 94602
HD 68256
HD 68257
HD 72945
HD 147722
BD+213825
IRC ID
0348356-373717
1055371+244457
0812129+173850
0812129+173850
0835509+063714
1624398-294215
1933467+213057
Contamination Source
HD 24072
HD 94601
HD 68257
HD 68256
HD 72946
HD 147723
2MASS J19334644+2130586a
a 2MASS J19334644+2130586 shows IR excess.
from the MIR detector onboard AKARI in the SAA regions. They
found that ionising particle hits affected the detector, leading to
changes in the output offset. Therefore the SAA does not signif-
icantly affect measurements of point sources in the AKARI/IRC
All-Sky Survey data, and therefore the measured flux densities
at 9 µm of HD 166191, HD 39415, and HD 145263 are reliable.
2.5. KnownYSOsandlate-typestars
Since we focus on debris disks in this paper, we need to dis-
tinguish debris disks from other types of excess emission such
as those often seen in young stellar objects (YSOs) or late-type
stars. We searched for the attributes of the 42 stars exhibiting
18 µm excess in the SIMBAD database2 and find that 12 and
6 stars were classified as YSOs and late-type stars in the liter-
ature, respectively. The known YSOs and late-type stars in the
18 µm excess sample are listed in Table 4. The spectral type
and luminosity classes in Table 4 were taken from the SIMBAD
database, which is supposed to contain the latest information.
For stars whose stellar information was revised after the Tycho-
2 Spectral Type Catalogue, the spectral type and luminosity class
listed in Table 4 may differ from those in the Tycho-2 Spectral
Type Catalogue. Thus stars other than luminosity class V have
been included. 24 main-sequence stars remain as candidates for
debris disks after excluding the known YSOs and late-type stars.
2 URL: http://simbad.u-strasbg.fr/simbad/.
3
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
r
e
b
m
u
N
25
20
15
10
5
0
A stars
-1
0
1
3
2
4
Ks-[L18W]
5
6
7
r
e
b
m
u
N
35
30
25
20
15
10
5
0
F stars
-1
0
1
3
2
4
Ks-[L18W]
5
6
7
r
e
b
m
u
N
25
20
15
10
5
0
G stars
-1
0
1
3
2
4
Ks-[L18W]
5
6
7
M stars
r
e
b
m
u
N
20
15
10
5
0
K stars
-1
0
1
3
2
4
Ks-[L18W]
5
6
7
r
e
b
m
u
N
5
4
3
2
1
0
-1
0
1
2
3
4
5
6
7
Ks-[L18W]
Fig. 2. Histogram of the KS−[18] colour of the stars detected at 18 µm for each spectral type star. The results of Gaussian fitting of
the histograms are indicated by the dashed lines. The three vertical dotted lines indicate the peak position µKS −[L18W] (middle), and
2σ positions (left and right) of the distribution of the KS−[18] colour.
Table 4. List of known YSOs and late-type stars detected at
18 µm.
Name
IRC ID
Fobs,9
(Jy)
HD 2326
HD 23937
HD 37258
HD 37357
HD 38238
HD 68695
HD 100453
HD 107439
HD 120806
HD 135344
HD 139614
HD 142666
HD 144432
HD 152404
HD 155555
HD 157045
HD 158643
HD 223075
0027064-063616
0348475-070053
0536592-060916
0537471-064230
0544187+000840
0811445-440508
1133055-541928
1221125-491240
1351516-034033
1515484-370915
1540463-422953
1556400-220140
1606579-274310
1654448-365318
1717255-665704
1727235-711715
1731250-235745
2346235+032912
99.943 ± 0.725
37.028 ± 0.144
1.749 ± 0.004
1.513 ± 0.001
1.221 ± 0.004
0.722 ± 0.010
5.518 ± 0.077
4.724 ± 0.422
58.500 ± 0.627
1.727 ± 0.063
2.386 ± 0.057
5.152 ± 0.355
2.032 ± 0.025
0.816 ± 0.025
80.030 ± 0.106
--
--
174.197 ± 4.056
Fobs,18
(Jy)
58.287 ± 0.641
13.536 ± 0.217
1.622 ± 0.046
1.671 ± 0.060
1.690 ± 0.059
1.275 ± 0.061
18.744 ± 0.248
3.556 ± 0.132
29.671 ± 0.085
3.285 ± 0.114
10.568 ± 0.118
6.578 ± 0.007
7.326 ± 0.259
2.870 ± 0.029
0.282 ± 0.027
85.198 ± 0.310
9.241 ± 0.111
44.205 ± 0.808
KS−[9]
(mag)
0.86
0.47
3.92
3.44
2.69
3.35
3.08
4.07
0.97
2.06
3.32
3.48
--
2.90
0.11
3.18
--
0.72
KS−[18]
(mag)
1.95
1.05
5.53
5.23
4.72
5.65
6.08
5.44
1.91
4.44
6.61
5.42
5.35
4.95
0.63
4.92
4.01
0.91
Spectral
Type
M7
M3
A2V
A0e
A7IIIe
A0V
A9Ve
G4Vp
M
A0V
A7Ve
A8Ve
A9/F0V
F5Ve
G7IV
M3III
A0V
CII
Stellar
Type
Ref.a
O-rich AGB
O-rich AGB
Herbig Ae
Herbig Ae
Herbig Ae
Herbig Ae
Herbig Ae
post-AGB
O-rich AGB
Herbig Ae
Herbig Ae
Herbig Ae
Herbig Ae
T Tauri
T Tauri
O-rich AGB
Herbig Ae
Carbon Star
8
9
2
2
3
4
2
6
8
2
1
2
2
2
7
10
1
5
a References. (1) Montesinos et al. (2009); (2) Acke et al. (2009); (3)
Hern´andez et al. (2005); (4) Manoj et al. (2006); (5) Abia et al. (2009);
(6) Gielen et al. (2008); (7) Schutz et al. (2009); (8) Winters et al.
(2003); (9) Sloan & Price (1998); (10) Josselin et al. (1998).
2.6. PhotosphericSEDFittingandExtinctionDetermination
The photospheric flux densities of the KS−[18]-selected candi-
dates were estimated from the Kurucz model (Kurucz 1992) fit-
ted to the 2MASS JHKS-band photometry of the stars taking
account of the extinction in the NIR.
We
by
Fitzpatrick & Massa (2009), which is a generalization of
the NIR reddening
adopt
curve
given
4
the analytic formula given by Pei (1992). For λ ≥ λ0, it has a
form of
E(λ − V)
E(B − V)
≡ RV Aλ
AV
− 1! =
0.349 + 2.087RV
1 + (λ/λ0)α
− RV ,
(1)
)
e
e
r
g
e
d
(
C
E
D
)
)
e
e
e
e
r
r
g
g
e
e
d
d
(
(
C
C
E
E
D
D
)
)
e
e
e
e
r
r
g
g
e
e
d
d
(
(
C
C
E
E
D
D
)
)
e
e
e
e
r
r
g
g
e
e
d
d
(
(
C
C
E
E
D
D
)
)
e
e
e
e
r
r
g
g
e
e
d
d
(
(
C
C
E
E
D
D
80
60
40
20
0
-20
-40
-60
-80
80
80
60
60
40
40
20
20
0
0
-20
-20
-40
-40
-60
-60
-80
-80
80
80
60
60
40
40
20
20
0
0
-20
-20
-40
-40
-60
-60
-80
-80
80
80
60
60
40
40
20
20
0
0
-20
-20
-40
-40
-60
-60
-80
-80
80
80
60
60
40
40
20
20
0
0
-20
-20
-40
-40
-60
-60
-80
-80
0
0
50
50
100
100
150
200
150
200
RA (degree)
RA (degree)
250
250
300
300
350
350
K stars (#=2122)
G stars (#=14013)
0
0
50
50
100
100
200
150
150
200
RA (degree)
RA (degree)
250
250
300
300
350
350
M stars (#=76)
K stars (#=2122)
0
0
50
50
100
100
150
200
150
200
RA (degree)
RA (degree)
250
250
300
300
350
350
Fig. 1. Spatial distributions of A-M dwarf stars from the Tycho-2
Spectral Type Catalogue (Wright et al. 2003).
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
A stars (#=18232)
where λ0 = 0.507 µm. Thus, the absolute extinction Aλ/AV is
given by
0
50
100
150
200
RA (degree)
250
300
350
F stars (#=29766)
A stars (#=18232)
0
0
50
50
100
100
150
200
150
200
RA (degree)
RA (degree)
250
250
300
300
350
350
Aλ
AV
=
0.349 + 2.087RV
1 + (λ/λ0)α
·
1
RV
.
(2)
The ratio of the total to selective extinction, RV ≡ A(V)/E(B−V),
and α are free parameters. NIR observations show that these pa-
rameters can vary from one sight line to another. In this work, we
use a curve defined by α = 2.05 and RV = 3.11. These parame-
ters provide a good representation of the mean curve determined
by Fitzpatrick & Massa (2009) and also of the extinction data
taken from the literature for the diffuse interstellar medium in
the Milky Way.
In our photospheric SED fitting, the scaling factor of lumi-
nosity with varying distance and extinction, and the extinction
AV are free parameters, while the effective temperature and the
surface gravity are inferred from the spectral type. The derived
photospheric flux densities at 9 and 18 µm for each star are listed
in Table 5.
G stars (#=14013)
F stars (#=29766)
2.7. SelectedDebrisDiskCandidates
The final catalogue contains 24 AKARI-identified debris disk
candidates showing 18 µm excesses. Among these, we iden-
tify 8 stars (HD 64145, HD 75809, HD 89125, HD 106797,
HD 113457, HD 165014, HD 175726, and HD 176137) are
new debris disk candidates, and the excess emission of 2 stars
(HD 15407 and HD 39415), which have never been confirmed
following their tentative discovery with IRAS (Oudmaijer et al.
1992), is confirmed by our AKARI survey. The SEDs of the 24
identified debris disk candidates are shown in Figure 3. We sum-
marise the observed and photospheric flux densities (Fobs and
F∗, respectively) and excess significance (χ ≡ (Fobs − F∗)/σ)
at 9 and 18 µm of all the candidates in Table 5, where σ is the
uncertainty of the excess level calculated from the errors of the
AKARI/IRC PSC photometry and in the estimation of the photo-
spheric emission from the 2MASS photometry for each source.
13 stars out of the 24 identified debris disk candidates also show
excess at 9 µm. It should be noted that since the uncertainty in
the flux density for a single source is calculated as deviation
of photometry of multiple detections in the AKARI/IRC PSC,
the uncertainty may possibly be too small in case the number
of detections is small, such as the cases of 18 µm photometry
of HD 165014 and HD 166191. The median value of the un-
certainty at 18 µm in our debris disk candidates is 0.022 Jy.
Therefore the 18 µm excess towards HD 165014 and HD 166191
will still be significant even if we assume the median values as
the photometric uncertainties for the two stars.
We list the stellar parameters (spectral type, stellar mass, lu-
minosity, age, and distance) of all the candidates in Table 6. The
spectral type of the stars is all taken from the SIMBAD database.
For HD 3003, HD 9672, HD 15407, HD 39060, HD 89125,
HD 98800, HD 106797, HD 109573, HD 110058, HD 145263,
and HD 181296, we take stellar masses and luminosities from
literatures. For the other stars whose stellar masses and lumi-
nosities are unavailable from the literature, we estimate them
from their spectral types according to Cox (2000).
5
Table 5. List of debris disk candidates with 18 µm excess.
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Name
IRC ID
HD 3003
HD 9672
HD 15407
HD 39060
HD 39415
HD 64145
HD 75809
HD 89125
HD 98800
HD 105209
HD 106797
HD 109573
HD 110058
HD 113457
HD 113766
HD 121617
HD 123356
HD 145263
HD 165014
HD 166191
HD 167905
HD 175726
HD 176137
HD 181296
0032440-630153
0134378-154035
0230506+553254
0547170-510359
0554415+443007
0753297+264556
0850454-381440
1017143+230621
1122052-244639
1206526-593529
1217062-654135
1236009-395211
1239461-491156
1305023-642630
1306357-460201
1357411-470035
1407340-210437
1610551-253121
1804432-205643
1810303-233401
1818182-232819
1856371+041553
1858451+020707
1922512-542526
Fobs,9
(Jy)
0.586 ± 0.010
0.368 ± 0.008
0.960 ± 0.031
3.216 ± 0.009
0.244 ± 0.017
0.762 ± 0.005
0.397 ± 0.012
0.975 ± 0.003
1.083 ± 0.028
0.392 ± 0.006
0.266 ± 0.017
0.360 ± 0.025
0.065 ± 0.016
0.146 ± 0.010
1.308 ± 0.065
0.103 ± 0.008
1.392 ± 0.046
0.236 ± 0.017
1.362 ± 0.043
1.816 ± 0.127
2.364 ± 0.007
0.433 ± 0.005
5.870 ± 1.112
0.622 ± 0.016
F∗,9
(Jy)
0.639
0.395
0.295
2.384
0.085
0.852
0.399
0.988
0.357
0.274
0.236
0.291
0.055
0.128
0.149
0.078
1.133
0.042
0.150
0.101
0.250
0.419
1.319
0.626
χ9
-1.7
-2.5
20.9
1.6
9.2
-0.8
-0.1
-0.1
25.0
14.9
1.6
2.6
0.6
1.7
17.8
3.0
4.7
11.4
28.0
13.5
231.8
1.2
4.1
-0.1
Excess
at 9 µm
Fobs,18
(Jy)
Yes
Yes
Yes
Yes
Marginal
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
0.238 ± 0.016
0.199 ± 0.014
0.497 ± 0.021
4.799 ± 0.042
0.274 ± 0.015
0.303 ± 0.041
0.152 ± 0.021
0.368 ± 0.025
6.153 ± 0.325
0.236 ± 0.013
0.148 ± 0.017
1.371 ± 0.022
0.084 ± 0.006
0.107 ± 0.025
1.428 ± 0.050
0.324 ± 0.027
0.567 ± 0.033
0.486 ± 0.013
0.959 ± 0.004
2.466 ± 0.002
1.771 ± 0.078
0.215 ± 0.034
4.995 ± 0.600
0.340 ± 0.021
F∗,18
(Jy)
0.137
0.084
0.064
0.511
0.018
0.183
0.088
0.214
0.077
0.060
0.050
0.062
0.012
0.027
0.032
0.017
0.248
0.009
0.033
0.022
0.055
0.090
0.299
0.134
χ18
5.9
8.2
20.6
35.1
17.1
2.5
3.0
3.1
18.7
13.5
5.7
59.4
12.0
3.2
27.9
11.4
9.5
36.7
224.1
1133.5
22.0
3.7
7.8
9.3
Excess
at 18 µm
Marginal
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
Yes
3. Additional Photometric Measurements of Debris
Disk Candidates
3.1. Subaru/COMICSandGemini/T-ReCSFollow-up
Observations
HD 15407, HD 165014, and HD 175726 were observed with
the COoled Mid-Infrared Camera and Spectrometer (COMICS;
Kataza et al. 2000; Okamoto et al. 2003; Sako et al. 2003)
mounted on the 8 m Subaru Telescope during 2007 June and
July. Imaging observations at the 8.8 µm (∆λ = 0.8 µm), 9.7 µm
(∆λ = 0.9 µm), 10.5 µm (∆λ = 1.0 µm), 11.7 µm (∆λ = 1.0 µm),
12.4 µm (∆λ = 1.2 µm), and 18.8 µm (∆λ = 0.9 µm) bands were
carried out. The pixel scale was 0.′′13 pixel−1. To cancel out the
background radiation, a secondary mirror chopping method was
used. We used standard stars from Cohen et al. (1999) as a flux
calibrator, and determined the reference point-spread functions
(PSFs) from the observations. We observed standard stars before
or after the observations of the target star in the same manner as
the target stars. The parameters of the COMICS observations are
summarised in Table A.1.
HD 106797, HD 105209, and HD 110058 were ob-
served with the Thermal-Region Camera Spectrograph (T-
ReCS; Telesco et al. 1998), mounted on the 8 m Gemini South
Telescope during 2007 June and July. Imaging observations at
the 8.8 µm (∆λ = 0.8 µm; Si2), 9.7 µm (∆λ = 0.9 µm; Si3),
10.4 µm (∆λ = 1.0 µm; Si4), 11.7 µm (∆λ = 1.1 µm; Si5),
12.3 µm (∆λ = 1.2 µm; Si6), and 18.3 µm (∆λ = 1.5 µm;
Qa) bands were carried out. The pixel scale was 0.′′09 pixel−1.
Procedures for the observations were basically the same as for
the COMICS observations. The parameters of the T-ReCS ob-
servations are summarised in Table A.2.
For the data reduction, we used our own reduction tools and
IRAF3. The standard chop-nod pair subtraction and the shift-
and-add method in units of 0.1 pixel were employed. We ap-
plied an air mass correction by estimating the difference be-
tween the target and standard stars in atmospheric absorption
using ATRAN (Lord 1992). The difference in the air masses be-
tween the target and standard stars was small (. 0.1) in general
and the correction factor in each band was less than 5 %.
The derived flux densities are summarised in Tables 7 and 8
for the COMICS and T-ReCS targets, respectively. All of them
are in good agreement with the flux densities of the AKARI ob-
servations taking account of the narrow band widths of filters
of the ground-based observations and possible dust features in
these spectral ranges (Fujiwara et al. 2009, 2010, 2012b).
3.2. IRAS,MSX,Spitzer and WISE Photometry
We searched for counterparts whose positional offsets were
smaller than 1′ in IRAS (Beichman et al. 1988) Point Source
Catalogue (PSC) and Faint Source Catalogue (FSC), and
MSX6C Infrared Point Source Catalogue (Egan et al. 2003) of
the Midcourse Space Experiment (MSX; Price et al. 2001) via
VizieR for each debris disk candidate. The results are shown
in Tables 9 and 10 for IRAS and MSX, respectively. We also
searched for Spitzer/MIPS data from published literatures and
found those of HD 3003 (Smith et al. 2006) and HD 181269
(Su et al. 2006; Rebull et al. 2008), as shown in Table 11.
Recent all-sky observations in the MIR by Wide-field
Infrared Survey Explorer (WISE) have provided the All-
Sky Source Catalogue (Cutri et al. 2012), which is similar to
AKARI/IRC PSC and is worth comparing. We searched for coun-
terparts of our MIR excess stars whose positional offsets were
smaller than 5′′ in the WISE All-Sky Source Catalogue for com-
parison. The results are shown in Table 12. All of our debris
3 IRAF is distributed by the National Optical Astronomy
Observatory, which is operated by the Association of Universities
for Research in Astronomy (AURA) under cooperative agreement with
the National Science Foundation.
6
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
10
1
0.1
0.01
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.001
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
HD3003
Kurucz Mmodel
Inner Disk (T=185K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD39415
Kurucz Mmodel
Inner Disk (T=321K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD98800
Kurucz Mmodel
Inner Disk (T=190K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD110058
Kurucz Mmodel
Inner Disk (T=104K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD123356
Kurucz Mmodel
Inner Disk (T=365K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD167905
Kurucz Mmodel
Inner Disk (T=467K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
HD9672
Kurucz Mmodel
Inner Disk (T= 73K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD64145
HD64145
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD105209
Kurucz Mmodel
Inner Disk (T=332K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD113457
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD145263
Kurucz Mmodel
Inner Disk (T=272K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD175726
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
100
HD15407
HD39060
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
Kurucz Mmodel
Inner Disk (T=553K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD75809
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD106797
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD113766
Kurucz Mmodel
Inner Disk (T=369K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD165014
Kurucz Mmodel
Inner Disk (T=488K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD176137
Kurucz Mmodel
Inner Disk (T=401K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
0.1
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
10
1
0.1
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
0.01
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
100
)
y
J
(
y
i
t
i
s
n
e
D
x
u
F
l
10
1
0.1
0.01
1
Kurucz Mmodel
Inner Disk (T=106K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD89125
Kurucz Mmodel
Inner Disk (T=170K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD109573
Kurucz Mmodel
Inner Disk (T=158K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD121617
Kurucz Mmodel
Inner Disk (T=173K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD166191
Kurucz Mmodel
Inner Disk (T=340K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
HD181296
Kurucz Mmodel
Inner Disk (T=123K)
Kurucz + Inner Disk
10
Wavelength (micron)
100
Fig. 3. NIR -- FIR SED of our debris disk candidates with 18 µm excess. The pluses, filled squares, open circles, crosses, open
diamonds, open triangles, and filled circles indicate the photometry with the 2MASS, AKARI/IRC, IRAS, MSX, WISE, ground-based
observations (Subaru/COMICS and Gemini/T-ReCS), and Spitzer/MIPS, repectively. We also plot contributions of the photosphere
(dashed lines) and those of inner disk component (dotted lines) fitted with AKARI/IRC, IRAS, and Spitzer/MIPS measurements by
assuming that the debris dust emits blackbody emission.
Table 8. Summary of Gemini/T-ReCS photometry.
Object
HD 105209
HD 106797
HD 110058
Si2 8.8
(mJy)
271 ± 27
201 ± 20
--
Si3 9.7
(mJy)
280 ± 28
165 ± 17
--
Si4 10.4
(mJy)
292 ± 29
159 ± 15
--
Si5 11.7
(mJy)
331 ± 33
180 ± 14
39 ± 4
Si6 12.3
(mJy)
299 ± 30
160 ± 13
--
Qa 18.8
(mJy)
238 ± 31
119 ± 59
--
7
Table 9. IRAS photometry of the debris disk candidates with 18 µm excess.
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Name
HD 3003
HD 9672
HD 15407
HD 15407
HD 39060
HD 39415
HD 64145
HD 75809
HD 89125
HD 98800
HD 105209
HD 109573
HD 110058
HD 113766
HD 121617
HD 123356
HD 145263
HD 166191
HD 167905
HD 176137
HD 181296
IRAS12
(Jy)
IRAS25
(Jy)
0.48 ± 0.06
0.33 ± 0.03
1.05 ± 0.06
1.05 ± 0.06
3.46 ± 0.21
0.30 ± 0.03
0.72 ± 0.07
0.35 ± 0.10
0.78 ± 0.10
1.98 ± 0.14
0.49 ± 0.04
< 0.43
< 0.38
1.59 ± 0.08
< 0.25
1.36 ± 0.10
0.43 ± 0.07
2.35 ± 0.75
4.14 ± 0.25
4.91 ± 0.79
0.54 ± 0.08
0.32 ± 0.04
4.06 ± 0.07
0.43 ± 0.04
0.43 ± 0.04
9.05 ± 0.45
0.27 ± 0.09
< 0.64
< 0.25
< 0.25
9.28 ± 0.74
0.37 ± 0.40
3.38 ± 0.30
0.22 ± 0.05
1.76 ± 0.12
0.72 ± 0.09
0.61 ± 0.07
0.52 ± 0.08
3.80 ± 0.57
2.67 ± 0.27
4.01 ± 0.76
0.48 ± 0.04
IRAS60
(Jy)
< 0.40
IRAS100
(Jy)
< 1.00
2.00 ± 0.20
1.91 ± 0.19
19.9 ± 1.4
11.3 ± 0.90
7.28 ± 0.80
4.46 ± 0.54
< 25.0
3.81 ± 0.42
< 0.40
< 0.40
< 0.40
< 0.40
< 0.41
< 0.40
< 1.78
7.36 ± 0.96
0.45 ± 0.06
0.66 ± 0.05
1.95 ± 0.18
7.18 ± 1.29
< 0.40
< 1.62
< 2.76
< 11.6
0.53 ± 0.04
< 1.14
< 1.14
< 1.23
< 1.03
< 5.28
< 1.00
< 1.04
< 1.00
< 1.45
< 1.00
< 3.49
< 208
< 59.2
< 127
< 1.00
IRAS ID
Catalogue
00304-6318
01321-1555
02272+5519
02272+5519
05460-5104
05510+4429
07504+2653
08488-3803
10144+2321
11195-2430
12043-5919
12333-3935
12369-4855
13037-4545
13545-4645
14047-2050
16078-2523
18074-2334
18152-2329
18562+0202
19188-5431
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
PSC
FSC
PSC
PSC
PSC
PSC
Table 10. MSX photometry of the debris disk candidates with 18 µm excess.
Band C (12.13 µm) Band D (14.65 µm)
Name
HD 105209
HD 106797
HD 113457
HD 165014
HD 166191
HD 167905
HD 175726
HD 176137
Band A (8.28 µm)
(Jy)
0.347 ± 0.018
0.272 ± 0.013
0.140 ± 0.010
1.077 ± 0.045
2.028 ± 0.081
2.772 ± 0.114
0.439 ± 0.020
5.497 ± 0.236
(Jy)
< 0.952
< 0.449
< 0.601
0.783 ± 0.074
1.863 ± 0.117
2.912 ± 0.163
< 0.573
6.102 ± 0.305
(Jy)
< 0.728
< 0.354
< 0.466
0.764 ± 0.065
1.953 ± 0.129
2.279 ± 0.148
< 0.450
4.517 ± 0.276
Band E (21.3 µm)
(Jy)
MSX ID
< 2.082
< 1.039
< 1.349
--
3.166 ± 0.212
1.487 ± 0.126
< 1.262
3.601 ± 0.230
G297.3128+02.7932
G299.4048-03.0570
G304.3984-01.6086
G009.0807+00.3009
G007.4400-02.1430
G008.3752-03.6697
G037.3180+00.7938
G035.6512-00.6602
disk candidates with 18 µm excesses have a counterpart in the
the WISE All-Sky Source Catalog.
4. Results and Discussion
4.1. DebrisDiskFrequency
The derived apparent debris disk frequency (the fraction of the
debris disk candidates in those stars detected at 18 µm in the
AKARI/IRC All-Sky Survey) in our sample is summarised in
Table 1. We find an overall apparent debris disk frequency of
2.8 ± 0.6% (24/856). The apparent debris disk frequency for A,
F, G, K, and M stars is estimated to be 5.6 ± 1.6% (11/196),
3.1 ± 1.0% (10/324), 1.2 ± 0.8% (2/173), 0.7 ± 0.7% (1/144),
0.0% (0/19), respectively. Since it is limited by the sensitivity of
the AKARI/IRC All-Sky Survey, the sample is biased towards
stars exhibiting MIR excess. To estimate the debris fraction
properly, we count only the excess sample whose photosphere
would be detectable at 18 µm with the AKARI/IRC All-Sky
Survey (i.e. F∗,18 & 90 mJy; HD 3003, HD 39060, HD 64145,
HD 75809, HD 89125, HD 123356, HD 175726, HD 176137,
and HD 181296) among 18 µm-detected sources with detectable
photosphere, and calculate the "real" debris disk frequency rela-
tive to those with F∗,18 & 90 mJy. We find an overall debris disk
frequency of 1.1±0.4% (9/856). The debris disk frequency for A,
F, G, K, and M stars is derived as 2.2 ± 1.1% (4/178), 1.0 ± 0.6%
(3/311), 1.2 ± 0.8% (2/169), 0.0% (0/144), 0.0% (0/19), respec-
tively, suggesting a possible tendency of an increase in the debris
disk frequency towards earlier type stars. The statistics of the de-
bris disk frequency is summarises in Table 1.
The excess rates at 24 µm by Spitzer/MIPS observa-
tions for A stars and solar-type FGK stars are 32% (52/160)
(Su et al. 2006) and 6% (5/82) (Beichman et al. 2006), respec-
tively. Therefore the debris disk frequency derived in this work
is much smaller than those estimated by Spitzer/MIPS obser-
vations both for A and FGK stars. It should be noted that
the uncertainty in Spitzer/MIPS pointed observations is smaller
than the present catalogue of the AKARI/IRC All-Sky Survey
and thus Spitzer/MIPS observations detect fainter excess than
this work. Our thresholds for 18 µm excess detection for the
AKARI/IRC sample in this study are shown in Table 2, which
can potentially detect a debris disk star with 18 µm excess of
Fobs,18/F∗,18 & 1.4. If we take a debris disk sample with 24 µm
excess from the Spitzer/MIPS sample with the same threshold
(Fobs,24/F∗,24 & 1.4) as set for the AKARI/IRC sample in this
work, the frequency becomes 14% (22/160) and 1% (1/82) for A
and FGK stars, respectively. The debris disk frequency of FGK
stars in this work is derived as 0.8 ± 0.3% (5/624), being consis-
tent with the Spitzer/MIPS work (Beichman et al. 2006). On the
other hand, the frequency of A stars in this work is still smaller
than that of the Spitzer/MIPS work (Su et al. 2006) even after
considering the difference in the excess detection threshold. The
difference might come from a selection effect of the input stars;
8
Table 12. WISE photometry of the debris disk candidates with 18 µm excess.
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Name
HD 3003
HD 9672
HD 15407
HD 39060
HD 39415
HD 64145
HD 75809
HD 89125
HD 98800
HD 105209
HD 106797
HD 109573
HD 110058
HD 113457
HD 113766
HD 121617
HD 123356
HD 145263
HD 165014
HD 166191
HD 167905
HD 175726
HD 176137
HD 181296
W1 (3.4 µm)
(Jy)
3.066 ± 0.198
2.026 ± 0.115
1.721 ± 0.093
10.603 ± 1.023
0.463 ± 0.013
4.182 ± 0.328
1.819 ± 0.096
5.611 ± 0.429
1.976 ± 0.116
1.086 ± 0.051
1.217 ± 0.048
2.028 ± 0.119
0.298 ± 0.007
0.697 ± 0.021
0.896 ± 0.039
0.435 ± 0.013
4.624 ± 0.358
0.241 ± 0.005
1.010 ± 0.036
0.787 ± 0.026
1.938 ± 0.086
2.467 ± 0.176
3.055 ± 0.203
3.184 ± 0.209
W2 (4.6 µm)
(Jy)
2.356 ± 0.084
1.311 ± 0.033
1.376 ± 0.040
10.808 ± 0.406
0.279 ± 0.005
3.010 ± 0.113
0.966 ± 0.023
4.530 ± 0.235
1.273 ± 0.038
0.715 ± 0.016
0.731 ± 0.014
1.143 ± 0.029
0.162 ± 0.003
0.384 ± 0.008
0.660 ± 0.015
0.239 ± 0.004
3.747 ± 0.169
0.153 ± 0.003
1.083 ± 0.026
0.879 ± 0.020
3.100 ± 0.125
1.593 ± 0.055
4.868 ± 0.239
2.369 ± 0.078
W3 (12 µm)
(Jy)
0.363 ± 0.005
0.232 ± 0.003
0.699 ± 0.010
2.698 ± 0.017
0.220 ± 0.003
0.416 ± 0.006
0.249 ± 0.003
0.522 ± 0.007
1.807 ± 0.020
0.301 ± 0.004
0.159 ± 0.002
0.278 ± 0.004
0.047 ± 0.001
0.104 ± 0.002
1.265 ± 0.011
0.076 ± 0.001
0.932 ± 0.012
0.292 ± 0.004
0.851 ± 0.011
2.157 ± 0.028
2.052 ± 0.025
0.236 ± 0.003
3.369 ± 0.044
0.405 ± 0.006
W4 (22 µm)
(Jy)
0.258 ± 0.005
0.268 ± 0.006
0.404 ± 0.009
7.972 ± 0.081
0.279 ± 0.006
0.128 ± 0.004
0.112 ± 0.003
0.139 ± 0.003
6.995 ± 0.065
0.212 ± 0.004
0.162 ± 0.003
2.714 ± 0.035
0.228 ± 0.005
0.099 ± 0.003
1.672 ± 0.015
0.588 ± 0.010
0.497 ± 0.009
0.468 ± 0.009
0.703 ± 0.012
3.271 ± 0.046
1.516 ± 0.027
0.106 ± 0.006
2.691 ± 0.050
0.418 ± 0.007
WISE ID
J003244.02-630154.0
J013437.83-154034.8
J023050.75+553253.3
J054717.01-510357.5
J055441.52+443007.5
J075329.80+264556.5
J085045.44-381439.9
J101714.20+230621.6
J112205.23-244639.4
J120652.71-593529.7
J121706.26-654134.7
J123600.95-395210.8
J123946.17-491155.6
J130501.99-642629.8
J130635.77-460202.0
J135741.10-470034.4
J140734.01-210437.5
J161055.09-253121.9
J180443.14-205644.6
J181030.32-233400.6
J181818.22-232819.7
J185637.17+041553.6
J185845.10+020706.9
J192251.21-542527.0
we select only dwarf A stars from the Tycho-2 Spectral Type
Catalogue, while the Su et al. (2006) sample includes a number
of very young star (age<10 Myr) and late B stars, which show
excess more frequently.
The frequency of debris disks around M stars in this work is
derived as 0.0% (0/19). Although the only known M-type debris
disk AU Mic is detected with IRC with 1.08 ± 0.01 and 0.25 ±
0.02 Jy at 9 and 18 µm, respectively, no MIR excess is found
towards the star. The non-detection of excess at 9 and 18 µm
around AU Mic is consistent with MIR observations with Spitzer
(e.g. Chen et al. 2005).
4.2. CharacteristicsofSelectedDebrisDiskCandidates
4.2.1. MIR colour, Inner Temperature, and Inner Radius
Estimate of the dust temperature is important to infer the radial
structures of debris disks. MIR observations are particularly use-
ful to trace the temperature of the inner hottest region of de-
bris disks. We derived the dust temperature from the 9 to 18 µm
colours of the observed excess emission for each star by assum-
ing that the debris dust emits blackbody emission. This temper-
ature is supposed to correspond to the dust temperature (Tin) at
the inner radius (Rin) of the debris disk.
In the case where excess emission at 9 µm was not detected
with AKARI/IRC, the dust temperature should be lower than
the temperature for those which show excess at 9 µm. For the
sources whose 9 µm excess was not detected with AKARI/IRC
and excess at the wavelengths of λ & 24 µm was reported
by IRAS or Spitzer observations, we derived the dust tempera-
ture from blackbody fitting to the observed excess emission at
λ & 24 µm. In the case that excesses at 9 µm and at & 24 µm
were not detected with AKARI/IRC, IRAS, and Spitzer, we set an
upper limit of Tin at 170 K, which is the highest possible tem-
perature of the coldest debris disk with 9 and 18 µm excess in
the sample (Tin = 158+11
−13 for HD 109573). We also derived the
inner radius of debris disk (Rin) as
Rin = r L∗
16πσSB
T −2
in ,
(3)
where L∗ and σSB is the stellar luminosity and the Stefan-
Boltzmann constant, respectively. The derived MIR colour, Tin,
and Rin are listed in Table 13.
The inner temperature Tin and inner radius Rin of A stars
range from . 100 K to ∼ 330 K and from ∼ 5 AU to > 20 AU,
respectively. The median value of the inner temperature among
A stars is . 170 K. On the other hand, Tin and Rin of solar-type
FGK stars ranges from . 170 K to ∼ 500 K and from < 1 AU to
& 5 AU, respectively. The median value of the inner temperature
among FGK stars is ∼ 350 K, much higher than that of A stars.
This suggests that the origin of the observed MIR excess is dif-
ferent between A and FGK stars: while the MIR excess emission
of A stars in our sample is typically contributed from the Wien-
side tail of the emission from the cool debris material, that of
FGK stars originates in the emission from the inner warm debris
material.
4.2.2. Lack of Cooler Material in Outer Region
The flux contributed from the inner dust for each star is shown
in Figure 3. The excess component of our samples except for
HD 109573, HD 121617, and HD 166191 can be fitted fairly
well only with the inner dust component of a single black-
body. This implies the lack of cooler material in the outer re-
gion around those stars and truncation of the disk at inner radii,
although FIR and sub-mm observations with high sensitivity are
required to accurately investigate the presence of the outer mate-
rial. Only one object (HD 166191) shows a rising SED towards
60 µm among our 13 FGK-type samples, while about half of
the samples of 11 A-type stars show FIR excess, suggesting that
solar-type stars in our sample tend to have less cooler material
in the outer region than A stars.
9
Table 13. Inner temperature and radius of debris disk candidates with 18 µm excess.
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Ldust/L∗
Cooler
Material
Yes
Yes
Yes
L−0.5
∗
t−1
age
(4)
fmax = 0.16 × 10−3 R7/3
dust M−5/6
∗
for the fixed model parameters (belt width: dr/r = 0.5; plan-
etesimal strength: Q∗
D = 200 J/kg; planetesimal eccentricity:
e = 0.05; and diameter of the largest planetesimal in cascade:
Dc = 2000 km). The values of fmax were calculated using the
parameters of the central star listed in Table 6 and the estimated
parameters for the inner disk component in Table 13.
The ratios of the observed fractional luminosities to the the-
oretical maximum fractional luminosities ( fobs/ fmax) are sum-
marised in Table 14. Some debris disks show fractional lumi-
nosities of fobs/ fmax ≫ 100, suggesting that these debris disks
cannot be accounted for by simple models of the steady-state
debris disks, and that transient events likely play a significant
role, even taking account of the uncertainties in the models
(Wyatt et al. 2007a).
One of the transient phenomena producing a large amount of
debris dust around a star is a dynamical instability that scatters
planetesimals inward from a more distant planetesimal belt. Dust
is released from unstable planetesimals following collisions and
sublimation. This is akin to the late heavy bombardment (LHB)
in the solar system, the cataclysmic event that occurred about
700 Myr after the initial formation of the solar system, as implied
by Moon's cratering record (e.g. Hartmann et al. 2000).
To develop a LHB-like event around a star, a newly born
planet would have to migrate and stir up planetesimals. Since
planet formation is expected to last for several hundred Myr
around solar-type stars, LHB-like events could be the cause of
dense warm debris disks around the young (age . Gyr) FGK sys-
tems of HD 113766, HD 145263, and HD 98800. On the other
hand, HD 89125, HD 15407, and HD 175726 are older than
one Gyr and planet formation is expected to have already been
finished. Therefore another dust production mechanism would
be required to account for the large fractional luminosities of
Tin
(K)
185+21
−21
73+4
−4
553+34
−31
106+5
−5
321+17
−17
(< 170)
(< 170)
(< 170)
190+3
−3
332+13
−13
(< 170)
158+11
−13
104+1
−1
(< 170)
369+13
−13
173+12
−14
365+38
−38
272+8
−9
488+12
−13
340+11
−12
467+15
−15
(< 170)
401+67
−61
123+5
−5
Rin
(AU)
10.4+2.8
−2.0
69.5+8.3
−7.0
0.5+0.1
−0.1
20.4+2.1
−1.8
1.6+0.2
−0.2
(> 16.9)
(> 5.6)
(> 4.0)
1.8+0.1
−0.1
5.8+0.5
−0.4
(> 17.5)
15.3+2.9
−1.9
22.7+0.4
−0.4
(> 22.2)
1.2+0.1
−0.1
21.5+3.9
−2.7
0.7+0.2
−0.1
2.1+0.1
−0.1
0.9+0.0
−0.0
1.4+0.1
−0.1
0.7+0.0
−0.0
(> 2.9)
1.0+0.4
−0.3
24.1+2.1
−1.8
written by
7 × 10−5
7 × 10−4
6 × 10−3
2 × 10−3
6 × 10−3
(> 9 × 10−5)
(> 2 × 10−4)
(> 3 × 10−4)
8 × 10−2
4 × 10−4
(> 2 × 10−4)
2 × 10−3
2 × 10−3
(> 3 × 10−4)
2 × 10−2
2 × 10−3
8 × 10−4
1 × 10−2
2 × 10−2
5 × 10−2
2 × 10−2
(> 6 × 10−4)
9 × 10−3
2 × 10−4
Name
HD 3003
HD 9672
HD 15407
HD 39060
HD 39415
HD 64145
HD 75809
HD 89125
HD 98800
HD 105209
HD 106797
HD 109573
HD 110058
HD 113457
HD 113766
HD 121617
HD 123356
HD 145263
HD 165014
HD 166191
HD 167905
HD 175726
HD 176137
HD 181296
Fexc,9
(Jy)
--
--
0.665
--
0.159
--
--
--
0.726
0.118
--
0.069
--
--
1.160
0.025
0.258
0.194
1.212
1.715
2.114
--
4.552
--
Fexc,18
(Jy)
0.101
0.115
0.434
4.288
0.256
0.120
0.064
0.154
6.076
0.177
0.097
1.309
0.072
0.080
1.396
0.308
0.319
0.477
0.927
2.445
1.717
0.125
4.696
0.206
Fexc,9/Fexc,18
--
--
1.53 ± 0.10
--
0.62 ± 0.08
--
--
--
0.12 ± 0.01
0.67 ± 0.06
--
0.05 ± 0.02
--
--
0.83 ± 0.06
0.08 ± 0.03
0.81 ± 0.17
0.41 ± 0.04
1.31 ± 0.05
0.70 ± 0.05
1.23 ± 0.06
--
0.97 ± 0.27
--
Rhee et al. (2007) suggest that the dust temperature in typi-
cal debris disks discovered with IRAS observations is less than
100 K and the peak of the excess flux density comes to around
60 − 100 µm. However, most of the debris disks detected by
AKARI/IRC in this work have abundant warm dust, while cooler
material in the outer regions is not conspicuous, suggesting that
we detect a sub-group of the debris disks, in which warm dust
grains are more abundant than in debris disks detected by IRAS.
Special mechanisms, such as dust trapping by the resonance per-
turbation of planets or generation of a large amount of warm
dust, might play significant roles in the inner region of disk in
these systems.
4.2.3. Fractional Luminosity of Inner Disk
Estimating the amount of dust is important to characterise the
debris disk. One of the observable indicators of the dust abun-
dance is its fractional luminosity, f = Ldust/L∗, i.e., the IR lumi-
nosity from the disk divided by the stellar luminosity. We derive
the fractional luminosities ( fobs) of the inner debris disk by in-
tegrating the intrinsic stellar emission and the observed excess
emission of the inner debris component. The derived fractional
luminosities are shown in Table 13. It should be noted that the
fractional luminosities for the inner disks with the cold dust com-
ponent should be regarded as upper limits, since we do not take
account of the contribution from the cold dust. The derived frac-
tional luminosities range from ∼ 10−4 to ∼ 10−1. The fractional
luminosity of our own zodiacal cloud is estimated to be of the or-
der ∼ 10−7 (Backman & Paresce 1993). Thus the debris disks in
our sample are typically by more than 1000 times brighter than
our own zodiacal cloud.
Wyatt et al. (2007a) developed a simple model for the
steady-state evolution of debris disks produced by collisions and
suggested that the maximum fractional luminosity ( fmax) can be
10
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Table 6. Stellar parameters of debris disk candidates with 18 µm
excess.
Name
HD 3003
HD 9672
HD 15407
HD 39060
HD 39415
HD 64145
HD 75809
HD 89125
HD 98800
HD 105209
HD 106797
HD 109573
HD 110058
HD 113457
HD 113766
HD 121617
HD 123356
HD 145263
HD 165014
HD 166191
HD 167905
HD 175726
HD 176137
HD 181296
Spectral
Type
A0V
A1V
F3V
A6V
F5V
A3V
F3V
F8Vbw
K5V
A1V
A0V
A0V
A0V
A0V
F4V
A1V
G1V
F0V
F2V
F4V
F3V
G5
F5
A0Vn
M∗
(M⊙)
2.9
2.7
1.4
1.8
1.4
2.5
1.4
0.98
0.67
2.9
2.9
2.9
2.9
2.9
1.4
2.9
1.05
1.6
1.6
1.4
1.4
1.01
1.4
2.9
L∗
(L⊙)
21.0
22.8
3.9
8.7
4.3
39.8
4.3
2.19
0.7
68.7
42.4
24.3
10.0
68.7
4.3
68.7
1.4
4.1
7.2
4.3
4.3
1.2
4.3
22.0
Age
(Myr)
50
20
80 -- 2100
12+8
−4
--
--
--
6457
10
--
10 -- 20
8
10
10 -- 20
10 -- 20
8 -- 10
--
--
--
--
--
--
4800 ± 3500
12+8
−4
d
(pc)
47
61
55
19
--
78
--
23
47
--
96
67
100
95
123
120
--
116
--
--
53
27
--
48
Ref.a
1
1
8,9
6
5
5
5
11
13
5
3,4
1
1
4,5
10
5
5
7
5
5
5
12
5
2
(2007b);
a References.
(1) Wyatt et al.
(2) Zuckerman & Webb
(2000); (3) Fujiwara et al. (2009); (4) de Zeeuw et al. (1999); (5)
Cox (2000); (6) Crifo et al. (1997); (7) Honda et al. (2004); (8)
Holmberg et al.
(10) Uzpen et al.
(2007); (11) Takeda (2007); (12) Bruntt (2009); (13) Wyatt et al.
(2007a).
(9) Melis et al.
(2010);
(2009);
Table 7. Summary of Subaru/COMICS photometry.
Object
HD 15407
HD 165014
HD 175726
N8.8
(mJy)
904 ± 90
927 ± 93
--
N11.7
(mJy)
644 ± 64
894 ± 89
214 ± 20
Q18.8
(mJy)
486 ± 73
838 ± 126
--
Table 11. Spitzer/MIPS photometry of the debris disk candidates
with 18 µm excess.
Name
HD 3003
HD 181296
MIPS24 MIPS70 MIPS160 Ref.a
(mJy)
224 ± 9
382 ± 7
(mJy)
62 ± 5
409 ± 42
1
2,3
(mJy)
--
68
a References.
(1) Smith et al.
(2006);
(2) Su et al.
(2006);
(3)
Rebull et al. (2008).
HD 89125, HD 15407, and HD 175726. It is worth noting that
recent FIR observations of HD 15407 by Herschel and AKARI
revealed the absence of cold dust around the star, not supporting
an LHB-like event as the origin of the bright debris disk around
the star (Fujiwara et al. 2012a).
4.2.4. Difference between A and FGK stars
To investigate the relationship between the debris disks and the
spectral type of the central stars, we plot the distributions of Tin,
Table 14. Comparison of observed ( fobs) and theoretical frac-
tional luminosities ( fmax) for our debris disk sample whose age
is available.
Name
HD 3003
HD 9672
HD 15407
HD 39060
HD 89125
HD 98800
HD 106797
HD 109573
HD 110058
HD 113457
HD 113766
HD 145263
HD 175726
HD 181296
fobs
7 × 10−5
7 × 10−4
6 × 10−3
2 × 10−3
(> 3 × 10−4)
8 × 10−2
(> 2 × 10−4)
2 × 10−3
2 × 10−3
2 × 10−2
1 × 10−2
2 × 10−4
fmax
7 × 10−5
2 × 10−3
(300 -- 1) ×10−8
3 × 10−3
(> 7 × 10−7)
6 × 10−5
(> 1 × 10−3)
5 × 10−5
2 × 10−3
7 × 10−6
4 × 10−5
1 × 10−3
(> 3 × 10−4)
(> 2 × 10−3)
(> 6 × 10−4)
(> (40 -- 1) ×10−7)
fobs/ fmax
1
0.4
20000 -- 600000
0.9
(> 500)
1000
(> 0.2)
40
0.7
(> 0.2)
3000
300
(> 200 -- 6000)
0.2
Rin, fractional luminosity and fobs/ fmax sorting by the spectral
types in Figure 4. It is seen that all of the debris disks inves-
tigated here change their characteristics appreciably around the
spectral type F0; Tin . 200 K for A stars while Tin & 300 K for
FGK stars; Rin & a few AU for A stars while Rin . a few AU for
FGK stars; the fractional luminosities for A stars are distributed
around 10−5 − 10−3 while those for FGK stars are in a range
10−3 − 10−1, and fobs/ fmax for A stars is distributed around unity
while that for FGK stars is & 100. In other words, the properties
of the inner disks of A and FGK stars are different from each
other. Most of our debris disk samples around FGK stars possess
warm dust without an appreciable amount of cooler dust, while
those debris disks with 18 µm excesses around A stars show lit-
tle excess at 9 µm. Our debris disk samples around FGK stars
may belong to a sub-group of the debris disks in which violent
dust supply is taking place in the inner region (. a few AU) of
the disk.
Radiation pressure is capable of removing dust grains and
changing the radial distribution of dust. When the radiation-
pressure-to-gravity ratio β = Frad/Fgrav becomes larger than
unity, the outward force dominates and the dust grains are blown
out. The ratio β can be approximated by
β = (0.4 µm/D)(2.7 g/cm3/ρ)(L∗/M∗),
(5)
where D is the grain size, ρ is the grain density, and L∗ and
M∗ are in units of L⊙ and M⊙ (Burns et al. 1979). For µm-sized
(D ∼ 1 µm) silicate (ρ ∼ 2.7 g/cm3) grains, β becomes about
unity around a F0 star (M∗ = 1.6M⊙ and L∗ = 4.1L⊙). Therefore
µm-sized grains may be blown out by radiation pressure in the
vicinity of A stars, while they can survive in FGK stars.
Morales et al. (2009) observed 52 main-sequence A and late
B type stars with the Spitzer/IRS that showed excess emission at
24 µm in the Spitzer/MIPS observations. They found no promi-
nent spectral features evident in any of the spectra, suggesting
that fine particles that show spectral features are absent around
early-type stars.
By taking the estimated β value and the absence of small
dust grains around early-type stars into account, it is suggested
that µm-sized grains are blown out by radiation pressure from A
stars and that the difference of the disk characteristics between A
and FGK stars seen in our sample may be driven by the radiation
pressure on grains.
11
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
e
p
y
T
l
a
r
t
c
e
p
S
A0
A0
A0
A0
A0
A0
A1
A1
A1
A3
A6
F0
F2
F3
F3
F3
F4
F4
F5
F5
F8
G1
G6
K5
(a)
HD105209
HD3003
HD181296
HD109573
HD106797
HD113457
HD110058
HD9672
HD121617
HD64145
HD39060
HD145263
HD165014
HD15407
HD167905
HD75809
HD113766
HD176137
HD166191
HD39415
HD89125
HD123356
HD175726
HD98800
e
p
y
T
l
a
r
t
c
e
p
S
A0
A0
A0
A0
A0
A0
A1
A1
A1
A3
A6
F0
F2
F3
F3
F3
F4
F4
F5
F5
F8
G1
G6
K5
(b)
HD3003
HD181296
HD109573
HD106797
HD113457
HD110058
HD9672
HD105209
HD121617
HD64145
HD39060
HD145263
HD165014
HD15407
HD167905
HD75809
HD113766
HD166191
HD176137
HD39415
HD89125
HD123356
HD175726
HD98800
e
p
y
T
l
a
r
t
c
e
p
S
A0
A0
A0
A0
A0
A0
A1
A1
A1
A3
A6
F0
F2
F3
F3
F3
F4
F4
F5
F5
F8
G1
G6
K5
HD3003
HD181296
(c)
HD109573
HD106797
HD113457
HD110058
HD9672
HD105209
HD121617
HD64145
HD39060
HD145263
HD165014
HD15407
HD167905
HD75809
HD113766
HD166191
HD176137
HD39415
HD89125
HD123356
HD175726
HD98800
e
p
y
T
l
a
r
t
c
e
p
S
A0
A0
A0
A0
A0
A0
A1
A6
F0
F3
F4
F8
G6
K5
HD3003
HD181296
(d)
HD109573
HD106797
HD113457
HD110058
HD9672
HD39060
HD145263
HD15407
HD113766
HD89125
HD175726
HD98800
100 200 300 400 500 600
0.2 0.5 1 2 5 10 20 50
10-5 10-4 10-3 10-2 10-1 100
10-1100101102103104105106
Tin (K)
Rin (AU)
Fractional Luminosity
fobs/fmax
Fig. 4. Distributions of Tin (a), Rin (b), fractional luminosity (c) and fobs/ fmax (d). Filled and open squares indicate the stars with and
without 9 µm excess, respectively. We set Tin at 170 K for the stars without 9 µm excess and photometric information at λ & 24 µm.
4.3. IndividualStar
Here we present notes about the individual debris disk candi-
dates with 18 µm excess discovered with AKARI or confirmed
in this work for the first time after the report by Oudmaijer et al.
(1992).
4.3.1. HD 15407
HD 15407 is an F3V star at d = 55 pc, whose IR excess was
originally suggested by Oudmaijer et al. (1992) and confirmed
with our present AKARI study. The SIMBAD database indi-
cates HD 15407 is a star in double system with the nearby K2V
star HIP 11692 (HD 15407B) with the apparent separation of
21.′′2. In this paper we refer the primary star (frequently called
as HD 15407A) simply as HD 15407. HD 15407 star may be
mature and the age is estimated as 2.1 Gyr by Holmberg et al.
(2009). On the other hand, a recent paper (Melis et al. 2010)
suggests a much younger age (∼ 80 Myr) based on the lithium
6710 Å absorption and the motion in the Galaxy. This star is
one of the most peculiar debris disk candidates in our sample
since the fractional luminosity is much larger than the steady-
state evolution model of planetesimals ( fobs/ fmax > 105). In ad-
dition, abundant silica (SiO2) dust, which is rare dust species in
the universe, is seen in the Spitzer/IRS spectrum (Fujiwara et al.
2012b; Olofsson et al. 2012). Special mechanisms to generate a
large amount of peculiar dust play a significant role around the
star.
4.3.2. HD 39415
HD 39415 is an F5V star, whose IR excess was originally
suggested by Oudmaijer et al. (1992) and confirmed with our
present AKARI study. No age information is available at present.
Holmes et al. (2003) report non-detection of sub-mm emission
at 870 and 1300 µm (3σ limits are 18.75 and 10.68 mJy at 870
and 1300 µm, respectively) with the Heinrich Hertz Telescope
at the Submillimeter Telescope Observatory, suggesting that the
amount of cool material around the star, if any, is quite small.
4.3.3. HD 64145
HD 64145 is an A3V star at d = 78 pc, whose 18 µm excess
is found in this AKARI sample. No age information is avail-
able at present. This star is reported as a binary system with
a secondary on a 582-day orbit (e.g. Abt 2005). Although no
excess at 9 µm is detected with the AKARI/IRC measurements,
the IRAS 12 µm flux density is slightly higher than the photo-
spheric level. Since the inner radius of the disk is derived as
Rin ∼ 17 AU and thus an apparent separation of 0.′′22 is ex-
pected, an extended disk structure might be resolved spatially in
the scattered light with a high contrast coronagraph with adap-
tive optics on 8 -- 10-m class ground-base telescopes. It should
be noted that this star is contained in the WISE All-Sky Source
Catalogue (Cutri et al. 2012) as J075329.80+264556.5. The flux
density at 22 µm (W4 band) is estimated as ∼ 130 mJy in the
WISE All-Sky Source Catalogue and is consistent with the ex-
pected photoshperic level, suggesting no significant excess at
22 µm. Photometries with narrow/intermediate filters at around
20 µm are required for examination of the inconsistency in the
flux density between AKARI and WISE.
4.3.4. HD 75809
HD 75809 is an F3V star, whose 18 µm excess is found in this
AKARI sample. No age information is available at present. No
9 µm excess is found with the AKARI observations, suggest-
ing that the temperature of the debris disk is at the lowest end
amongst our samples.
4.3.5. HD 89125
HD 89125 is a nearby F8V star at d = 23 pc, whose 18 µm ex-
cess is found in this AKARI sample. This star is mature and the
age is estimated as 6.5 Gyr (Takeda 2007). The fractional lumi-
12
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
nosity of the star is much larger than the steady-state evolution
model of planetesimals ( fobs/ fmax > 500), suggesting that tran-
sient dust production events might play a significant role around
the star. It should be noted that the star is reported as being metal
poor ([Fe/H]= −0.39) and one of the most metal-poor debris sys-
tems among the known sample (Beichman et al. 2006). This star
is contained in the WISE All-Sky Source Catalogue (Cutri et al.
2012) as J101714.20+230621.6. The flux density at 22 µm (W4
band) is estimated as ∼ 140 mJy in the WISE All-Sky Source
Catalogue and is consistent with the expected photoshperic level,
suggesting no significant excess at 22 µm. Photometry with nar-
row/intermediate filters at around 20 µm is required for exami-
nation of the inconsistency.
4.3.6. HD 106797
HD 106797 is an A0V star at d = 99 pc, whose 18 µm excess
is found in this AKARI sample. de Zeeuw et al. (1999) classified
HD 106797 as a member of Lower Centaurus Crux group, whose
age is estimated as 10 −20 Myr. Fujiwara et al. (2009) suggested
the presence of narrow spectral features between 11 and 12 µm,
which is attributable to (sub-)µm-sized crystalline silicates based
on the multi-band photometric data collected in the follow-up
observations with Gemini/T-ReCS.
4.3.7. HD 113457
HD 113457 is an A0V star at d = 95 pc, whose 18 µm excess
is found in this AKARI sample. de Zeeuw et al. (1999) classifies
HD 113457 as a member of Lower Centaurus Crux group, whose
age is estimated as 10−20 Myr. Since the inner radius of the disk
is estimated to be Rin ∼ 22 AU and thus an apparent separation
of 0.′′23 is expected, an extended disk structure might be resolved
spatially in the scattered light with a high contrast coronagraph
with adaptive optics.
4.3.8. HD 165014
HD 165014 is an F2V star, whose 9 and 18 µm excesses are
found in this AKARI sample. No age information is available
at present. This star was also detected at 8 -- 14 µm with MSX
and possible excesses at the wavelengths have been suggested
by Clarke et al. (2005). Strong dust features attributable to crys-
talline enstatite (MgSiO3) are seen in its Spitzer/IRS spectrum
(Fujiwara et al. 2010) while crystalline forsterite (Mg2SiO4) fea-
tures, which is typically more abundant around young stars, are
not seen. Possible formation of enstatite dust from differentiated
parent bodies is suggested according to the solar system analog.
Special mechanisms to generate a large amount of crystalline
enstatite dust must play a role around the star (Fujiwara et al.
2010).
4.3.9. HD 175726
HD 175726 is a nearby G6V star at d = 27 pc, whose 18 µm ex-
cess is found in this AKARI sample. The stellar age is estimated
as 4.8 ± 3.5 Gyr by Bruntt (2009). The fractional luminosity
is larger than the steady-state evolution model of planetesimals
( fobs/ fmax > 103), suggesting that some transient dust produc-
tion events play a role around the star. Since the stellar age is in
a range during which LHB took place in the solar system, the
bright debris disk around HD 175726 might be produced by a
LHB-like phenomenon. It should be noted that this star is con-
tained in the WISE All-Sky Source Catalogue (Cutri et al. 2012)
as J185637.17+041553.6. Although the flux density at 22 µm
(W4 band) is estimated as ∼ 110 mJy in the WISE Catalogue
and suggests excess at the wavelength, it is considerably lower
than the AKARI measurement at 18 µm. The discrepancy may
suggest temporal variability of MIR excess towards HD 175726
over a few years. A recent study reveals a dramatic decrease in
the infrared excess seen in one of the warm dust debris can-
didates, the young Sun-like star TYC 8241-2652-1, in several
years, which suggests removal of warm debris from the system
in a very short time scale (Melis et al. 2012). Monitoring obser-
vations of HD 175726 around 20 µm could reveal the nature and
evolution of its debris system. The star is a CoRoT asteroseismic
target and the presence of weak solar-like oscillations is reported
(Mosser et al. 2009) while no hint of planets orbiting the star is
so far suggested.
4.3.10. HD 176137
HD 176137 is an F5 star, whose 18 µm excess is found in this
AKARI sample. No age information is available at present. This
star is located in the direction towards a highly obscured re-
gion in the Galaxy, whose interstellar extinction is estimated as
AV = 27.1 by "Galactic Dust Reddening and Extinction" at the
IRSA4. The extinction towards the star derived from our fitting
procedure of photosphere is Av = 13.0, much smaller than the es-
timated extinction, suggesting that the star is located near to us.
HD 176137 is a double star system with a separation of 0.′′5 -- 1′′
(at the epoch of 1910-2009) as resolved by speckle interferom-
etry (Douglass et al. 2000). The nature of the nearby star is not
known enough. Ground-based MIR observations of the star with
high spatial resolution are needed for further discussion on the
possible confusion in the excess from the nearby star.
5. Summary
We report the results of a systematic survey of warm (T &
150 K) debris disks based on photometric measurements at
18 µm take from the AKARI/IRC All-Sky Survey data. We have
found 24 debris disk candidates with bright MIR excess emis-
sion above the stellar photospheric emission out of 856 sources
that were detected at 18 µm. Among them 8 stars are newly dis-
covered in this work. We found that 13 stars of the 24 debris
disk candidates also show excess emission at 9 µm. The over-
all apparent debris disk frequency is derived as 2.8 ± 0.6% for
our sample. We have identified a tendency for the frequency to
increase towards earlier type stars.
The temperature, radius, and fractional luminosity of the in-
ner disk component of the candidates are derived from the 9 to
18 µm flux densities of the excess emission sources. We find
that A stars and solar-type FGK stars have different characteris-
tics of their debris disks; most of those around FGK stars pos-
sess warm dust without any cool dust component, while the de-
bris disks around A stars show little evidence for excess at 9 µm
and appear to have lower temperature (T . 200 K) debris dust.
In addition, considering only the objects for which the age in-
formation is available, we find that the fractional luminosities
of the inner debris disks around FGK stars cannot be explained
by steady-state evolutionary models of debris disks that are sus-
tained by collisions of planetesimals. On the other hand, the de-
bris disks we have detected around A stars are found to fit well
with the steady-state model. We propose that the debris disks
4 URL: http://irsa.ipac.caltech.edu/applications/DUST/.
13
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
around FGK stars belong to a sub-group of debris disks, in which
violent dust supply and processing are taking place in the inner
region (. a few AU) of the disk. The suggested difference in the
debris disk characteristics between A and solar-type FGK stars
can be explored by the next version of the AKARI/IRC PSC and
the data from WISE, which will dramatically improve the sensi-
tivity of MIR photometry and should allow us to study a larger
population of these very interesting objects.
Acknowledgements. This research is based on observations with the AKARI,
a JAXA project with the participation of ESA and on data collected at the
Subaru Telescope and the Gemini Observatory. It has made use of the SIMBAD
database and the VizieR catalogue access tool, CDS, Strasbourg, France and
the data products from 2MASS. H.F. and S.T. was financially supported by the
Japan Society for the Promotion of Science. This research was supported by
KAKENHI (07J02823 and 23103002).
References
Abia, C., de Laverny, P., Recio-Blanco, A., et al. 2009, Publications of the
Astronomical Society of Australia, 26, 351
Abt, H. A. 2005, ApJ, 629, 507
Acke, B., Min, M., van den Ancker, M. E., et al. 2009, A&A, 502, L17
Aumann, H. H., Beichman, C. A., Gillett, F. C., et al. 1984, ApJ, 278, L23
Backman, D. E. & Paresce, F. 1993, in Protostars and Planets III, ed. E. H. Levy
& J. I. Lunine, 1253 -- 1304
Bastian, U. & Roser, S. 1993, Catalog of Positions and Proper Motions: South
(Heidelberg: Astron. Rechen Inst.), ed. S. Bastian, U. & Roser
Beichman, C. A., Bryden, G., Stapelfeldt, K. R., et al. 2006, ApJ, 652, 1674
Beichman, C. A., Neugebauer, G., Habing, H. J., Clegg, P. E., & Chester, T. J.,
eds. 1988, Infrared astronomical satellite (IRAS) catalogs and atlases. Volume
1: Explanatory supplement, Vol. 1
Bruntt, H. 2009, A&A, 506, 235
Burns, J. A., Lamy, P. L., & Soter, S. 1979, Icarus, 40, 1
Chen, C. H., Patten, B. M., Werner, M. W., et al. 2005, ApJ, 634, 1372
Clarke, A. J., Oudmaijer, R. D., & Lumsden, S. L. 2005, MNRAS, 363, 1111
Cohen, M., Walker, R. G., Carter, B., et al. 1999, AJ, 117, 1864
Cox, A. N. 2000, Allen's astrophysical quantities, ed. A. N. Cox
Crifo, F., Vidal-Madjar, A., Lallement, R., Ferlet, R., & Gerbaldi, M. 1997,
A&A, 320, L29
Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003, 2MASS All Sky Catalog
of point sources., ed. Cutri, R. M., Skrutskie, M. F., van Dyk, S., Beichman,
C. A., Carpenter, J. M., Chester, T., Cambresy, L., Evans, T., Fowler, J., Gizis,
J., Howard, E., Huchra, J., Jarrett, T., Kopan, E. L., Kirkpatrick, J. D., Light,
R. M., Marsh, K. A., McCallon, H., Schneider, S., Stiening, R., Sykes, M.,
Weinberg, M., Wheaton, W. A., Wheelock, S., & Zacarias, N.
Cutri, R. M., Wright, E. L., Conrow, T., et al. 2012, Explanatory Supplement to
the WISE All-Sky Data Release Products, Tech. rep.
de Zeeuw, P. T., Hoogerwerf, R., de Bruijne, J. H. J., Brown, A. G. A., & Blaauw,
A. 1999, AJ, 117, 354
Douglass, G. G., Mason, B. D., Rafferty, T. J., Holdenried, E. R., & Germain,
M. E. 2000, AJ, 119, 3071
Egan, M. P., Price, S. D., Kraemer, K. E., et al. 2003, VizieR Online Data
Catalog, 5114, 0
Eiroa, C., Marshall, J. P., Mora, A., et al. 2011, A&A, 536, L4
Fitzpatrick, E. L. & Massa, D. 2009, ApJ, 699, 1209
Fricke, W., Schwan, H., Corbin, T., et al. 1991, Veroeffentlichungen des
Astronomischen Rechen-Instituts Heidelberg, 33, 1
Fricke, W., Schwan, H., Lederle, T., et al. 1988, Veroeffentlichungen des
Astronomischen Rechen-Instituts Heidelberg, 32, 1
Fujiwara, H., Onaka, T., Ishihara, D., et al. 2010, ApJ, 714, L152
Fujiwara, H., Onaka, T., Takita, S., et al. 2012a, ApJ, 759, L18
Fujiwara, H., Onaka, T., Yamashita, T., et al. 2012b, ApJ, 749, L29
Fujiwara, H., Yamashita, T., Ishihara, D., et al. 2009, ApJ, 695, L88
Gielen, C., van Winckel, H., Min, M., Waters, L. B. F. M., & Lloyd Evans, T.
2008, A&A, 490, 725
Hartmann, W. K., Ryder, G., Dones, L., & Grinspoon, D. 2000, The Time-
Dependent Intense Bombardment of the Primordial Earth/Moon System, ed.
R. K. . e. a. Canup, R. M., 493 -- 512
Hern´andez, J., Calvet, N., Hartmann, L., et al. 2005, AJ, 129, 856
Holmberg, J., Nordstrom, B., & Andersen, J. 2009, A&A, 501, 941
Holmes, E. K., Butner, H. M., Fajardo-Acosta, S. B., & Rebull, L. M. 2003, AJ,
125, 3334
Honda, M., Kataza, H., Okamoto, Y. K., et al. 2004, ApJ, 610, L49
Houk, N. 1978, Michigan Catalog of Two-dimensional Spectral Types for HD
Stars, Vol. 2 (Ann Arbor: Univ. Michigan Dept. Astron.), ed. N. Houk
14
Houk, N. 1982, Michigan Catalog of Two-dimensional Spectral Types for HD
Stars, Vol. 3 (Ann Arbor: Univ. Michigan Dept. Astron.), ed. N. Houk
Houk, N. & Cowley, A. P. 1975, Michigan Catalog of Two-dimensional Spectral
Types for HD Stars, Vol. 1 (Ann Arbor: Univ. Michigan Dept. Astron.), ed.
A. P. Houk, N. & Cowley
Houk, N. & Smith-Moore, M. 1988, Michigan Catalog of Two-dimensional
Spectral Types for HD Stars, Vol. 4 (Ann Arbor: Univ. Michigan Dept.
Astron.), ed. J. Warren, W. H.
Houk, N. & Swift, C. 1999, Michigan Catalog of Two-dimensional Spectral
Types for HD Stars, Vol. 5 (Ann Arbor: Univ. Michigan Dept. Astron.), ed.
C. Houk, N. & Swift
Ishihara, D., Onaka, T., Kataza, H., et al. 2010, A&A, 514, A1
Jaschek, C., Conde, H., & de Sierra, A. C. 1964, Observatory Astronomical La
Plata Series Astronomies, 28, 1
Josselin, E., Loup, C., Omont, A., et al. 1998, A&AS, 129, 45
Kataza, H., Okamoto, Y., Takubo, S., et al. 2000, in Presented at the Society
of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol. 4008,
Society of Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, ed. M. Iye & A. F. Moorwood, 1144 -- 1152
Kennedy, P. M. 1996, VizieR Online Data Catalog, 3078, 0
Kurucz, R. L. 1992, in IAU Symposium, Vol. 149, The Stellar Populations of
Galaxies, ed. B. Barbuy & A. Renzini, 225 -- +
Lecavelier Des Etangs, A., Vidal-Madjar, A., & Ferlet, R. 1996, A&A, 307, 542
Liseau, R., Eiroa, C., Fedele, D., et al. 2010, A&A, 518, L132
Lohne, T., Augereau, J.-C., Ertel, S., et al. 2012, A&A, 537, A110
Lord, S. D. 1992, A new software tool for computing Earth's atmospheric trans-
mission of near- and far-infrared radiation, Tech. rep.
Manoj, P., Bhatt, H. C., Maheswar, G., & Muneer, S. 2006, ApJ, 653, 657
Melis, C., Zuckerman, B., Rhee, J. H., & Song, I. 2010, ApJ, 717, L57
Melis, C., Zuckerman, B., Rhee, J. H., et al. 2012, Nature, 487, 74
Meyer, M. R., Carpenter, J. M., Mamajek, E. E., et al. 2008, ApJ, 673, L181
Montesinos, B., Eiroa, C., Mora, A., & Mer´ın, B. 2009, A&A, 495, 901
Morales, F. Y., Werner, M. W., Bryden, G., et al. 2009, ApJ, 699, 1067
Mosser, B., Michel, E., Appourchaux, T., et al. 2009, A&A, 506, 33
Mouri, A., Kaneda, H., Ishihara, D., et al. 2011, PASP, 123, 561
Murakami, H., Baba, H., Barthel, P., et al. 2007, PASJ, 59, 369
Okamoto, Y. K., Kataza, H., Yamashita, T., et al. 2003, in Presented at the
Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol.
4841, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, ed. M. Iye & A. F. M. Moorwood, 169 -- 180
Olofsson, J., Juh´asz, A., Henning, T., et al. 2012, A&A, 542, A90
Onaka, T., Matsuhara, H., Wada, T., et al. 2007, PASJ, 59, 401
Oudmaijer, R. D., van der Veen, W. E. C. J., Waters, L. B. F. M., et al. 1992,
A&AS, 96, 625
Pei, Y. C. 1992, ApJ, 395, 130
Pilbratt, G. L., Riedinger, J. R., Passvogel, T., et al. 2010, A&A, 518, L1
Price, S. D., Egan, M. P., Carey, S. J., Mizuno, D. R., & Kuchar, T. A. 2001, AJ,
121, 2819
Rebull, L. M., Stapelfeldt, K. R., Werner, M. W., et al. 2008, ApJ, 681, 1484
Rhee, J. H., Song, I., Zuckerman, B., & McElwain, M. 2007, ApJ, 660, 1556
Roeser, S. & Bastian, U. 1988, A&AS, 74, 449
Sako, S., Okamoto, Y. K., Kataza, H., et al. 2003, PASP, 115, 1407
Schutz, O., Meeus, G., Sterzik, M. F., & Peeters, E. 2009, A&A, 507, 261
Sloan, G. C. & Price, S. D. 1998, ApJS, 119, 141
Smith, P. S., Hines, D. C., Low, F. J., et al. 2006, ApJ, 644, L125
Su, K. Y. L., Rieke, G. H., Stansberry, J. A., et al. 2006, ApJ, 653, 675
Takeda, Y. 2007, PASJ, 59, 335
Telesco, C. M., Pina, R. K., Hanna, K. T., et al. 1998, in Presented at the
Society of Photo-Optical Instrumentation Engineers (SPIE) Conference, Vol.
3354, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, ed. A. M. Fowler, 534 -- 544
Thompson, M. A., Smith, D. J. B., Stevens, J. A., et al. 2010, A&A, 518, L134
Uzpen, B., Kobulnicky, H. A., Monson, A. J., et al. 2007, ApJ, 658, 1264
Winters, J. M., Le Bertre, T., Jeong, K. S., Nyman, L., & Epchtein, N. 2003,
A&A, 409, 715
Wright, C. O., Egan, M. P., Kraemer, K. E., & Price, S. D. 2003, AJ, 125, 359
Wyatt, M. C. 2008, ARA&A, 46, 339
Wyatt, M. C., Smith, R., Greaves, J. S., et al. 2007a, ApJ, 658, 569
Wyatt, M. C., Smith, R., Su, K. Y. L., et al. 2007b, ApJ, 663, 365
Yamamura, I., Makiuti, S., Ikeda, N., et al. 2010, VizieR Online Data Catalog,
2298, 0
Zuckerman, B. & Webb, R. A. 2000, ApJ, 535, 959
Table A.1. Log of Subaru/COMICS observations.
Fujiwara et al.: AKARI/IRC 18 µm Survey of Warm Debris Disks
Object
HD 15407
5 Lac
HD 15407
HD 15407
γ Aql
γ Aql
γ Aql
HD 165014
HD 165014
HD 165014
ǫ Sco
ǫ Sco
HD 175726
γ Aql
Filter
(µm)
18.8
18.8
11.7
8.8
8.8
11.7
18.8
18.8
11.7
8.8
11.7
8.8
11.7
11.7
Date
(UT)
16 July 2008
16 July 2008
17 July 2008
17 July 2008
17 July 2008
17 July 2008
16 July 2008
16 July 2008
17 July 2008
17 July 2008
17 July 2008
17 July 2008
17 July 2008
17 July 2008
Table A.2. Log of Gemini/T-ReCS observations.
Object
HD 105209
HD 110458
HD 105209
HD 105209
HD 105209
HD 105209
HD 105209
HD 110458
HD 110458
HD 110458
HD 110458
HD 110458
HD 106797
HD 110458
HD 106797
HD 106797
HD 106797
HD 106797
HD 106797
HD 110458
HD 110458
HD 110458
HD 110458
HD 110458
HD 110058
HD 110458
HD 110058
HD 110458
Filter
(µm)
18.3
18.3
12.3
11.7
10.4
9.7
8.8
8.8
9.7
10.4
11.7
12.3
18.3
18.3
12.3
11.7
10.4
9.7
8.8
8.8
9.7
10.4
11.7
12.3
11.7
11.7
18.3
18.3
Date
(UT)
10 June 2008
10 June 2008
11 June 2008
11 June 2008
11 June 2008
11 June 2008
11 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
10 June 2008
10 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
12 June 2008
26 June 2008
26 June 2008
Integ.
(s)
2172
116
116
116
116
116
116
58
58
58
58
58
811
116
174
116
58
174
58
58
58
58
58
58
869
58
260
116
Integ. Comment
(s)
2400
100
300
100
20
20
400
20
200
100
10
50
300
20
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Comment
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Standard
Appendix A: Logs of Subaru/COMICS and
Gemini/T-ReCS Follow-up Observations
Here we show the logs of follow-up MIR observations of some
debris disk candidates by Subaru/COMICS (Table A.1) and
Gemini/T-ReCS (Table A.2).
15
|
1109.0582 | 2 | 1109 | 2011-10-26T19:45:07 | The Optical and Near-Infrared Transmission Spectrum of the Super-Earth GJ1214b: Further Evidence for a Metal-Rich Atmosphere | [
"astro-ph.EP"
] | We present an investigation of the transmission spectrum of the 6.5 M_earth planet GJ1214b based on new ground-based observations of transits of the planet in the optical and near-infrared, and on previously published data. Observations with the VLT+FORS and Magellan+MMIRS using the technique of multi-object spectroscopy with wide slits yielded new measurements of the planet's transmission spectrum from 0.61 to 0.85 micron, and in the J, H, and K atmospheric windows. We also present a new measurement based on narrow-band photometry centered at 2.09 micron with the VLT+HAWKI. We combined these data with results from a re-analysis of previously published FORS data from 0.78 to 1.00 micron using an improved data reduction algorithm, and previously reported values based on Spitzer data at 3.6 and 4.5 micron. All of the data are consistent with a featureless transmission spectrum for the planet. Our K-band data are inconsistent with the detection of spectral features at these wavelengths reported by Croll and collaborators at the level of 4.1 sigma. The planet's atmosphere must either have at least 70% water by mass or optically thick high-altitude clouds or haze to be consistent with the data. | astro-ph.EP | astro-ph |
ApJ in press
Preprint typeset using LATEX style emulateapj v. 5/2/11
THE OPTICAL AND NEAR-INFRARED TRANSMISSION SPECTRUM OF THE SUPER-EARTH GJ 1214B:
FURTHER EVIDENCE FOR A METAL-RICH ATMOSPHERE
Jacob L. Bean1,8, Jean-Michel D´esert1, Petr Kabath2, Brian Stalder1, Sara Seager3, Eliza Miller-Ricci
Kempton4,8, Zachory K. Berta1, Derek Homeier5, Shane Walsh6, & Andreas Seifahrt7
ApJ in press
ABSTRACT
We present an investigation of the transmission spectrum of the 6.5 M⊕ planet GJ 1214b based
on new ground-based observations of transits of the planet in the optical and near-infrared, and
on previously published data. Observations with the VLT + FORS and Magellan + MMIRS using
the technique of multi-object spectroscopy with wide slits yielded new measurements of the planet's
transmission spectrum from 0.61 to 0.85 µm, and in the J, H, and K atmospheric windows. We
also present a new measurement based on narrow-band photometry centered at 2.09 µm with the
VLT + HAWKI. We combined these data with results from a re-analysis of previously published FORS
data from 0.78 to 1.00 µm using an improved data reduction algorithm, and previously reported values
based on Spitzer data at 3.6 and 4.5 µm. All of the data are consistent with a featureless transmission
spectrum for the planet. Our K-band data are inconsistent with the detection of spectral features at
these wavelengths reported by Croll and collaborators at the level of 4.1 σ. The planet's atmosphere
must either have at least 70% H2O by mass or optically thick high-altitude clouds or haze to be
consistent with the data.
Subject headings: planets and satellites: atmospheres -- planets and satellites: individual: GJ 1214b
-- techniques: photometric
1. INTRODUCTION
Recent results from the Kepler mission indicate that
exoplanets intermediate in size between the terrestrial
and the ice giant planets in the Solar System are common
at small orbital separations from main sequence stars
(Borucki et al. 2011; Howard et al. 2011). However, de-
spite the wealth of knowledge of the sizes and masses
for these planets that will be obtained from the Kepler
data and follow-up efforts, there will still be a substan-
tial gap in our understanding of these objects. This is
because knowledge of only the mass and radius of plan-
ets in this regime is not sufficient to determine their bulk
composition due to degeneracies that exist in theoreti-
cal models (Adams et al. 2008; Rogers & Seager 2010a).
Specifically, most planets with masses and radii in the
intermediate-size range can be modeled equally well by
substantially different combinations of mass and compo-
sition for the interior and surrounding atmosphere.
Fortunately,
there is a way to break the degen-
[email protected]
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden
Street, Cambridge, MA 02138, USA
2 European Southern Observatory, Alonso de Cordova 3107,
Casilla 19001, Santiago, Chile
3 Department of Earth, Atmospheric, and Planetary Sciences
and Department of Physics, Massachusetts Institute of Technol-
ogy, Cambridge, MA 02139, USA
4 Department of Astronomy and Astrophysics, University of
California, 1156 High Street, Santa Cruz, CA 95064, USA
5 Centre de Recherche Astrophysique de Lyon, UMR 5574,
CNRS, Universit´e de Lyon, ´Ecole Normale Sup´erieure de Lyon,
46 All´ee d'Italie, F-69364 Lyon Cedex 07, France
6 Magellan Fellow, Australian Astronomical Observatory and
Curtin Institute of Radio Astronomy, Curtin University, GPO
Box U1987, Perth, WA 6845, Australia.
7 Department of Physics, University of California, One Shields
Avenue, Davis, CA 95616, USA
8 Sagan Fellow
eracies between different models for a planet when
a unique solution for the bulk composition can not
be found. The method is to obtain constraints on
the planet's atmospheric composition (Miller-Ricci et al.
2009; Miller-Ricci & Fortney 2010). This information
provides a boundary condition to further constrain both
rigorous physical models and also cosmogonical argu-
ments about likely origin and evolutionary scenarios.
Therefore, detailed characterization of representative ob-
jects to further constrain their bulk compositions is an
important complement to the statistical study of the oc-
currence rate of populations that is being enabled by
Kepler.
The transiting planet GJ 1214b (Charbonneau et al.
2009) is one object that is potentially a key to unlocking
the mystery of the bulk compositions of intermediate-
size planets using atmospheric studies.
Its determined
mass (6.5 M⊕) and radius (2.65 R⊕) put it firmly in the
degenerate region of parameter space for interior and at-
mosphere mass and compositions. The only thing that
can be said with a high degree of confidence based on
just the mass and radius of the planet is that it must
have a substantial atmosphere because it is too large to
be composed solely of solid material (Rogers & Seager
2010b; Nettelmann et al. 2011).
Rogers & Seager (2010b) identified three possible ori-
gins for the gas layer on GJ 1214b and they presented dis-
tinct interior structure models based on these scenarios
that are consistent with its measured mass and radius.
The model scenarios are (1) a "Mini-Neptune" planet
that is composed mainly of solid rock and ice and that has
a significant primordial atmosphere accreted from the
protoplanetary nebula and with a composition roughly
similar to that of the Sun (i.e., hydrogen-dominated),
(2) a "Water World" planet composed primarily of water
ice and having a secondary gas envelope formed by sub-
2
Bean et al.
limation and dominated by water vapor, and (3) a true
"Super-Earth" planet composed of purely rocky material
with a secondary atmosphere formed by outgassing and
probably composed mainly of hydrogen. Distinguish-
ing between these competing models for GJ 1214b would
have important consequences because they each imply
a very different formation and evolutionary history for
the planet. Constraining the formation and evolution-
ary history for this planet would be an important step
in our quest to obtain a general understanding of the
intermediate-size planets for which this object appears
to be an archetype.
The most interesting characteristic of GJ 1214b is that
it orbits a very small M dwarf, and thus the system has
a planet-to-star radius ratio comparable to a Jupiter-size
planet transiting a Sun-like star. This quality makes it
the most feasible known intermediate-size planet for at-
mospheric studies using transit spectroscopy techniques.
The three models proposed for the planet would also
exhibit very different transmission spectrum features.
Mini-Neptune and true Super-Earth planets with their
hydrogen-dominated, and thus large scale height, atmo-
spheres would exhibit relatively large spectral features in
transmission owing to absorption by trace gases like wa-
ter and methane, and scattering by molecular hydrogen
and cloud or haze particles. On the other hand, a Water
World planet with a primarily water vapor atmosphere,
and thus small scale height, would exhibit relatively small
spectral features in transmission.
We recently obtained measurements of GJ 1214b's
transmission spectrum in the red optical (0.78 -- 1.00 µm)
with the FORS instrument on the VLT using a new
ground-based technique to constrain the planet's atmo-
spheric composition (Bean et al. 2010). This remark-
ably featureless spectrum is consistent with a model
for a metal-rich atmosphere with a small scale height,
and inconsistent with a model for cloud-free hydrogen-
dominated atmospheres, which would have a large scale
height. Clouds or haze in a hydrogen-dominated at-
mosphere could conceivably also yield a flat spectrum
consistent with the observations. Subsequently, we pre-
sented broadband photometric measurements obtained
at 3.6 and 4.5 µm with Spitzer that provided addi-
tional constraints on the composition of the atmosphere
(D´esert et al. 2011). These data were also consistent
with a featureless transmission spectrum when analyzed
in isolation and in combination with the FORS data, but
the overall interpretation of the flat spectrum remained
uncertain due to the possibility of clouds or haze pro-
viding a gray opacity source, and non-equilibrium abun-
dances of methane, which is the primary expected opac-
ity source at 3.6 µm.
In contrast
to the results showing a featureless
spectrum, Croll et al. (2011a) presented ground-based
near-infrared photometry observations that indicated
GJ 1214b's transmission spectrum has a large feature
in the Ks-band (bandpass approximately 2.0 -- 2.3 µm)
based on the measurement of a substantially deeper tran-
sit in that band as compared to the observations at other
wavelengths. The main interpretation of the observations
in this case is unambiguous; the large scale height of a
hydrogen-dominated atmosphere is required for such a
feature to be observed.
Crossfield et al. (2011) presented high-resolution spec-
troscopy of GJ 1214b between 2.1 -- 2.4 µm. They de-
tected no spectral
features, but could not comment
on the absolute depth of the transit at these wave-
lengths due to the purely differential nature of their data.
Crossfield et al. (2011) came to similar conclusions as
D´esert et al. (2011) did based on the Spitzer data be-
cause methane is also the expected source of wavelength-
dependent opacity in the their observational window.
The observational results showing a featureless trans-
mission spectrum for GJ 1214b and the result from
Croll et al. (2011a) indicating large spectral features in
the near-infrared are not strictly in conflict because
they either were obtained at different wavelengths [i.e.,
the Bean et al. (2010) and D´esert et al. (2011) studies]
or have different sensitivities [i.e., the Crossfield et al.
(2011) investigation]. Also, there is a plausible qualita-
tive explanation for all the observations: a hyrodgen-
dominated atmosphere depleted in methane and with
clouds or haze opaque at optical wavelengths. However,
there is some tension because, for all the plausibility of
the qualitative model, it is not supported by physical cal-
culations at this point, and it has to be finely tuned to
match the observations.
Given the outstanding issues in our understanding of
GJ 1214b's atmosphere, we were motivated to conduct
further observational studies of its transmission spectrum
to improve our understanding of this important world. In
particular, we aimed to make independent measurements
of the planet's transmission spectrum in the K-band re-
gion using a different technique than Croll et al. (2011a)
used, and also to make measurements further in the blue
part of the optical to search for the signature of scatter-
ing from cloud, haze, or gas particles. We present here
the results of these investigations. In §2 we describe our
observations and data reduction. In §3 we describe our
analysis of the data. We compare our new measurements
of the planet's transmission spectrum to previous mea-
surements and theoretical models in §4. We conclude in
§5 with a summary of the results.
2. DATA
2.1. Magellan + MMIRS
2.1.1. Observations
We observed transits of GJ 1214b on 2011 May 15 and
18 using the MMIRS instrument (McLeod et al. 2004)
on the Magellan (Clay) telescope at Las Campanas Ob-
servatory. We used the same multi-object spectroscopy
approach for the MMIRS observations as was used for
the FORS observations of GJ 1214b that we presented
previously (Bean et al. 2010). We gathered time-series
spectra for GJ 1214b and three other reference stars of
similar brightness within 6′ using slits with lengths of 30′′
and widths of 12′′. The off-axis guide and wavefront sen-
sor of the instrument, which is sensitive to wavelengths
between 0.6 and 0.9 µm, was used to feed corrections to
the telescope control and active optics systems. We used
an HK grism as the dispersive element and an HK filter
to isolate the first order spectra. Complete spectra from
1.23 to 2.48 µm with a dispersion of 6.6 A pixel−1 were
obtained for all the objects.
As described below, only the data obtained on the sec-
ond night are reliable. The observations for this transit
spanned 2.63 h from UT 06:15 to 08:53. Exposure times
3
Fig. 1. -- An example spectrum extracted from the MMIRS data
for GJ 1214. The broad bandpasses in the J, H, and K atmospheric
windowns are indicated by the grey boxes.
were 25 s, and the overhead per exposure including the
read and reset time of the detector was 14 s. A total of
239 exposures were obtained, 79 of which were in tran-
sit. The observations began when the field was at an
airmass of 1.20, and by the end the field had set to an
airmass of 1.58. Conditions were clear and the seeing
varied between 0.5 and 0.9′′ as estimated by the width
of the spatial profiles of the obtained spectra.
2.1.2. Data reduction
The MMIRS data were recorded using the "up-the-
ramp" sampling mode of the detector in which non-
destructive reads were obtained every 5 s and stored in
a data cube. In the first step of the data reduction, we
subtracted the first of these reads from the others on
a pixel-by-pixel basis in each frame to remove the bias
and persistence from previous exposures. The estimated
dark current of the detector is 0.06 e− s −1 pixel−1, which
is negligible for this experiment. After subtracting the
first read in the data cube for an exposure, we then fit a
slope to the up-the-ramp samples for each pixel to deter-
mine the count rate. We arrived at the final estimate of
the total pixel counts by multiplying the count rate by
the total exposure time.
The reduction of the MMIRS data cubes gave reason-
able results for the second data set, but we noticed a re-
peating pattern of discontinuities in the residuals when
fitting the up-the-ramp samples for the first data set.
The origin of this effect is likely a problem in the detec-
tor electronics. We chose not to include the first data
set in the final analysis because of this problem; in what
follows we focus only on the analysis of the data set ob-
tained on the 18th.
After the data processing to collapse the data cubes,
we applied spectroscopic flat-field corrections, subtracted
the background, and extracted one-dimensional spectra
for the objects using the same algorithms as for our previ-
ous multi-object spectroscopy observations (Bean et al.
2010). We established the wavelength scale of the ex-
tracted stellar spectra based on spectra of an Ar emission
lamp obtained following the transit observation using a
copy of the science mask with the slit widths set to 0.5′′.
We adopted a third-order polynomial form for the wave-
length solution. The wavelength calibration was done
independently for each of the four slits, and the maxi-
Fig. 2. -- The MMIRS K-band raw time-series data for GJ 1214
(filled circles) and the composite reference (open diamonds).
mum dispersion from the fits was 0.3 A. After applying
the wavelength solution to the data, we noticed offsets
between the positions of telluric absorption lines in the
spectra extracted from the first image. These offsets are
consistent with a misalignment in the rotation angle be-
tween the science mask and the calibration mask. We de-
termined and applied offset corrections to the zero point
of the wavelength scale of each object. The maximal
correction size was 38 A, which is equivalent to approxi-
mately six pixels.
The positions of the spectra on the detector varied in
both the spatial and spectral directions during the time-
series due to imperfect guiding and instrument flexure,
and this led to apparent shifts in the extracted 1d spec-
tra relative to each other. We cross-correlated the spec-
tra for each object relative to the first spectrum in the
time-series, and then shifted them by this amount to cor-
rect for the effect of the drift. The position of the spec-
tra drifted by 2.5 pixels over the course of the first 25
exposures, and then there was a sudden shift of approx-
imately 3 pixels. The position of data after the large
jump remained relatively stable with a rms of 0.15 pix-
els and only a small long-term drift totaling 0.13 pixels.
We removed the first 30 points from the time-series to
avoid problems caused by the large instability at the be-
ginning. There is still sufficient out-of-transit baseline
(45 minutes) with these points removed, and the model
fits are substantially better for the trimmed light curves.
An example spectrum of GJ 1214 is shown in Figure 1.
The signal-to-noise ratio (SNR) in the GJ 1214 spectra
estimated from photon counting statistics reaches up-
wards of 475 pixel−1 in the H-band atmospheric window
where the maximum number of counts were recorded.
We estimate that the resolution of the stellar spectra is
approximately 50 A (7.5 pixels). This estimate is based
on the intrinsic resolving power of the grism (R ≡ λ/∆λ
≈ 560 for a 1.0′′ slit), the atmospheric seeing, errors in
the correction for the motion of the stars on the slits dur-
ing the observations, and the uncertainty in the absolute
wavelength scale.
2.1.3. Extraction of the photometric time-series
4
Bean et al.
Fig. 3. -- Normalized residuals (circles) from the MMIRS broadband K-band light curve fits without the airmass (left panel) and the
spatial spectrum position (right panel) decorrelations as a function of those parameters. The decorrelation fits are indicated by the solid
lines.
We created broadband photometric time-series from
the extracted spectra in the J, H, and K atmospheric
windows, and four spectroscopic time-series in the K-
band of width 0.1 µm by summing the obtained spectra
within a bandpass. As discussed in §3, both the broad-
band and spectroscopic data provide insight to the com-
position of the planet's atmosphere. We do not feel that
the J- and H-band data are of sufficient quality to justify
subdividing them into spectroscopic channels.
The definitions of the utilized bands and the deter-
mined planet-to-star radius ratios (Rp/R⋆) from the light
curve modeling (see §3) are given in Table 1. The limits
of the broad bandpasses were chosen to match standard
near-infrared filter curves1 as closely as possible to avoid
regions with significant telluric water absorption and to
enable quick comparison with other results, in partic-
ular the Ks-band measurement of Croll et al. (2011a).
However, our data do not include the entire range of the
standard J-band, and the combination of the grism and
order-blocking filter we used has substantially different
transmission properties than broadband filters. The ex-
act locations of the bandpass edges were also set so that
the spectra were not cut in the middle of strong absorp-
tion lines. Therefore, appropriately weighted integration
of the atmospheric models for the planet are necessary for
robust comparison of results even at similar wavelengths.
The time-series data for GJ 1214 in each bandpass were
corrected by dividing out the sum of the data for the ref-
erence stars.
In the subsequent light curve modeling,
we found that including all of the reference stars in this
step yielded a worse correction than when one of them
was excluded. A possible explanation for this is that
the spectra for the problematic reference star fell along
the boundary between different read-out channels in the
detector. The motion of the spectra during the time-
series resulted in the spectra for this object moving back
and forth across the boundary and this could have intro-
duced additional noise in the data. For the final analysis,
Photometric Bandpasses & Transit Depths for the
MMIRS Data
TABLE 1
Band
Wavelength (µm)
Rp/R⋆
a
Broadband analysis
J
H
K
Spectroscopic analysis
channel 1
channel 2
channel 3
channel 4
1.25 -- 1.33
1.50 -- 1.77
1.98 -- 2.31
0.1158 ± 0.0024
0.1146 ± 0.0014
0.1158 ± 0.0006
1.98 -- 2.08
2.08 -- 2.18
2.18 -- 2.28
2.28 -- 2.38
0.1156 ± 0.0007
0.1163 ± 0.0007
0.1158 ± 0.0006
0.1163 ± 0.0011
a From an analysis with the transit parameters a/R⋆ and i fixed
to 14.97 and 88.94◦, respectively.
we excluded the data for the bad reference star, and we
used only the data for the other two reference stars to
correct the GJ 1214 time-series. The estimated photon-
limited uncertainties in the photometry for the reference
star were propagated through to the final uncertainties
in the GJ 1214 data. An illustration of the reference star
correction to the broadband K-band data is shown in
Figure 2.
After correcting the time-series for GJ 1214 using the
reference star data, we identified additional systematics
correlated with the airmass of the observations and the
positions of the spectra in the spatial dimension on the
detector. We fit for linear decorrelations against these
two parameters simultaneously with the light curve mod-
eling described in §3 to remove these effects (two free
parameters in addition to the normalization). Adding
higher-order terms and decorrelating against airmass and
spatial pixel position for the individual stellar fluxes (i.e.,
GJ 1214 and the reference stars) did not improve the
results. The correlations between the reference star-
corrected light curve residuals and the resulting linear
fits for these two parameters are shown in Figure 3.
1 See http://www.cfht.hawaii.edu/Instruments/Filters/wircam.html
for the properties of the filters used by Croll et al. (2011a).
The MMIRS broadband light curves for GJ 1214 after
all corrections and decorrelations exhibit residuals from
5
Fig. 4. -- An example spectrum extracted from the FORS blue
data for GJ 1214
a best-fit model with rms of 1532, 887, and 706 ppm in
the J-, H-, and K-bands, respectively. These values are
factors of 3.9, 6.0, and 3.4 higher than expected from the
photon-limited uncertainties. The H-band data exhibit
an especially larger-than-expected scatter of unknown
origin. We experimented with changing the bandpasses
to investigate this issue. Shifting the bandpasses and also
using smaller and larger windows did not significantly
affect the results. We speculate that perhaps the trans-
parency of the Earth's atmosphere varies on different spa-
tial scales among the near-infrared atmospheric windows;
and that this scale, along with the number, color, and
spatial distribution of reference stars, sets the fundamen-
tal limit when working in the . 1000 ppm photon-limited
regime in the near-infrared. When binned over 60 s,
the rms of the K-band residuals is 443 ppm. These are
the highest-precision ground-based near-infrared transit
light curve data that we are aware of. The rms of the
residuals for the K-band spectroscopic channels range
from 923 ppm to 1552 ppm, and the scaling from the ex-
pected levels of noise range from 2.3 to 2.8.
2.2. VLT + FORS
We observed one transit of GJ 1214b on 2011 July 3
using the multi-object spectroscopy mode of the FORS
instrument (Appenzeller et al. 1998) on UT1 of the VLT
facility at Paranal Observatory. The instrument setup
and data reduction were very similar to the previous
FORS observations described in Bean et al. (2010) and
Berta et al. (2011), and we limit our discussion here to
the unique aspects of the current study.
We used the 600RI grism with the GG435 filter to ob-
tain spectra over the range 610 -- 850 nm for GJ 1214 and
five reference stars. Observations began at UT 02:04 and
continued for 3.12 h to UT 05:11. The field was at an air-
mass of 1.20 at the beginning, rose to an airmass of 1.15,
and then set to an airmass of 1.33 by the end. The ex-
posure times were 18 s, which gave a total duty cycle of
55 s including 37 s of overhead. A total of 194 exposures
were obtained, 57 of which were during the transit.
Our approach for the data reduction, spectral extrac-
tion, and wavelength calibration were the same as for
the previous FORS observations (Bean et al. 2010), with
one exception. During the wavelength calibration step,
we identified a few lamp lines in the FORS atlas that
must have inaccurate assigned wavelengths because they
Fig. 5. -- The broadband FORS blue raw time-series data for
GJ 1214 (filled circles) and the composite reference (open dia-
monds).
were significantly discrepant from a low-order polyno-
mial dispersion model. When we did not include these
lines in the solution, it became clear that a third-order
polynomial model for the dispersion was justified instead
of the linear model we had used previously and that is
also advocated in the instrument manual. The residuals
from the third-order dispersion fit with the bad lines ig-
nored are typically 0.005 nm, whereas the residuals are
typically 0.5 nm from a linear model with the bad lines
included.
This new insight to the wavelength scale of FORS
grism data prompted us to re-analyze the previous data
for the two transits of GJ 1214b from Bean et al. (2010),
and we present the new results below. We believe that
the resolution of the data can be reliably increased now
that the wavelength solution has been significantly im-
proved, and we re-bin the old spectroscopic data to 10 nm
instead of the 20 nm used before. The reduction and
correction of the data are otherwise unchanged. The
basic properties of the data remain the same after the
re-analysis, and the conclusions from Bean et al. (2010)
are strengthened.
In the rest of the paper we refer to
the new FORS data as the FORS "blue" data, and the
re-analyzed old data as the FORS "red" data.
An example spectrum of GJ 1214 extracted from the
FORS blue data is shown in Figure 4. The signal-to-
noise ratio in the GJ 1214 spectra estimated from pho-
ton counting statistics reaches upwards of 300 pixel−1 at
900 nm, but only 50 pixel−1 at 600 nm due to GJ 1214's
extremely red spectral energy distribution.
We created both a broadband and spectroscopic light
curves from the FORS blue data by summing over the full
wavelength range and over twelve channels with widths
of 20 nm, which is the lowest we feel that we can go
without systematic errors given the quality of the data.
We combined the data for four of the reference stars and
divided the composite from the GJ 1214 time-series to
correct it. The fifth reference star was found to yield a
worse correction to the photometry of GJ 1214 and was
not included in our final analysis. An illustration of the
reference star correction to the broadband data is shown
in Figure 5.
6
Bean et al.
TABLE 2
Mid-transit Times
Instrument Epocha
Tc (BJDTDB)b
O - C (s)c
MMIRS
FORS
HAWKI
464
493
517
2455699.832866 ± 6.9E-5
-7.6 ± 5.9
2455745.664729 ± 6.0E-5 +3.1 ± 5.2
2455783.59440 ± 1.1E-4
-0.4 ± 9.5
a Integer number of orbital periods since the transit on 2009 May
15.
b BJDTDB is the the Barycentric Julian Date in the Barycentric
Dynamical Time standard (Eastman et al. 2010).
c Residuals form the ephemeris given in §3.
in the field-of-view within the linearity range of the de-
tector. To improve the duty cycle, six separate exposures
were averaged on the chip to produce a single recorded
image. The total time to integrate the exposures, read
out the data, and reset the detector was 28 s. A total
of 337 images were obtained, 110 of which were during
transit. The maximum counts for GJ 1214, which was
the brightest star in the field-of-view, ranged from 9,000
to 25,000 e−. This is well below the the 1% non-linearity
level for the HAWKI detectors (approximately 60,000 e−)
and no non-linearity corrections were applied to the data.
The seeing reported by the observatory atmospheric sta-
tion varied from 0.9 to 3.2′′ during the observations, but
we recall that the guide probe of the telescope never reg-
istered more than 1.6′′ (this information is not included
in the image headers).
After bias subtraction and flat fielding the HAWKI
images, we measured aperture photometry for GJ 1214
and 11 other stars. The photometric aperture and sky
annulus sizes were optimized to give the lowest dispersion
in the light curve for GJ 1214. We settled on a stellar
aperture radius of 26 pixels, and a sky annulus inner
radius of 31 pixels and outer radius of 41 pixels for the
final analysis.
We corrected the time-series photometry of GJ 1214
by dividing out the sum of a number of the reference
stars. We experimented with different combinations of
the reference stars for the correction, and we ultimately
used a composite of four of them for the final analysis. A
plot of the raw time-series for GJ 1214 and the composite
reference is shown in Figure 6.
After the reference star correction, we noticed a slow
and smooth variation in the corrected GJ 1214 relative
flux values with time superimposed on the expected tran-
sit light curve morphology. This trend is potentially re-
lated to the large color difference between GJ 1214 and
the reference stars, although the bandpass of the utilized
filter is relatively narrow. Similar trends are regularly
seen in ground-based near-infrared transit photometry
(e.g., Croll et al. 2010a,b, 2011a,b; Sada et al. 2010), so
this probably arises from a property of Earth's atmo-
sphere. We found that the trend in our data was best
modeled by a second order polynomial with time, and
we fit for such a function simultaneously with the transit
modeling (see §3, two free parameters in addition to the
normalization). Similar results, but with higher disper-
sion in the light curve residuals and larger determined
parameter uncertainties, are obtained when decorrelat-
ing the data with a linear trend as a function of airmass.
The data from three of the images were found to deviate
by more than five times the standard deviation of the
Fig. 6. -- The HAWKI raw time-series data for GJ 1214 (filled
circles) and the composite reference (open diamonds).
As with our previous study of FORS data, we identified
a slow and smooth variation in the corrected relative flux
values with time superimposed on the expected transit
light curve morphology. This effect is probably due to the
smudges on the FORS atmospheric dispersion corrector
described by Moehler et al. (2010). We fit for a second-
order polynomial as a function of time simultaneously
with the transit modeling (see §3) to account for this
(two free parameters in addition to the normalization).
The FORS blue broadband light curve for GJ 1214 af-
ter all corrections and decorrelations exhibits residuals
from a best-fit model with a rms of 457 ppm. This is
a factor of 2.5 larger than expected from the estimated
photon noise. The rms of the residuals for the spectro-
scopic channels range from 665 ppm to 1785 ppm, and
the scaling from the expected levels of noise range from
1.2 to 1.4.
2.3. VLT + HAWKI
We observed one transit of GJ 1214b on 2011 August 10
using the standard imaging mode of the HAWKI instru-
ment (Kissler-Patig et al. 2008) on UT4 of the VLT facil-
ity at Paranal Observatory. The observations were done
using the NB2090 filter (λc = 2095 nm, width = 20 nm).
In contrast to previously reported HAWKI observa-
tions of exoplanet transits (e.g., Gillon et al. 2009;
Anderson et al. 2010; Gibson et al. 2010), we chose to
maintain a single pointing throughout the observations
instead of using a dither pattern. This "staring" ap-
proach has been used for observations with a similar in-
strument to obtain the highest-precision ground-based
transit light curves in the near-infrared up to now (ap-
proximately 700 ppm minute−1; e.g., Croll et al. 2010a,b,
2011b). We reach a similar level of precision in our data
using the staring approach, and this suggests that the
single pointing method is indeed the best technique to
use for exoplanet transit observations with HAWKI.
Observations began at UT 01:03 and continued for
2.68 h to UT 03:44. The field was at an airmass of 1.15 at
the beginning, and it set to an airmass of 1.64 by the end.
The exposure times were 1.6762 s, which is the minimum
possible for the normal mode of HAWKI. The telescope
was defocussed slightly to keep the counts for all the stars
7
Fig. 7. -- Upper panels Normalized broadband (left panel) and K-band spectroscopic (right panel) light curves from the MMIRS data
(circles). The best-fit models are shown as the solid lines. Bottom panels Residuals from the fits (circles).
residuals from a best-fit model. These discrepant points
were removed for the final analysis.
The HAWKI light curve for GJ 1214 after all correc-
tions, decorrelations, and trimming of discrepant points
exhibits residuals from a best-fit model with a rms of
1259 ppm. The rms of the residuals is 732 ppm when the
data are binned to 60 s. The dispersion of the per-image
residuals in the un-binned data is a factor of 1.3 larger
than expected from the estimated photon noise.
3. ANALYSIS
We fit the broadband and spectroscopic light curves
for GJ 1214 with a transit model multiplied by a nor-
malization factor and systematic decorrelation functions
to investigate the transmission spectrum of its transiting
planet. We used the exact analytic formulas including
quadratic limb darkening given by Mandel & Agol (2002)
for the transit model. The transit model was parameter-
ized by the planet-to-star radius ratio (Rp/R⋆), system
scale (a/R⋆), orbital inclination of the planet (i), mid-
transit time of the planet (Tc), and the stellar quadratic
limb darkening coefficients (γ1 and γ2). The orbital pe-
riod of the planet was fixed to the revised value given
below, and the eccentricity was fixed to zero. The nor-
malization factor for each data set was always a free pa-
rameter. The decorrelation functions described in §2 re-
quire two additional free parameters for each wavelength
bin.
We estimated limb darkening coefficients appropriate
for transits of GJ 1214b using spherically symmetric stel-
lar model atmospheres calculated with the PHOENIX
code (Hauschildt et al. 1999). We adopted the stellar
parameters for the baseline model Tef f = 3026 K, [M/H]
= 0.0, log g = 5.0, and R⋆ = 0.21 R⊙. We also computed
models at the baseline effective temperature ± 130 K. For
the MMIRS and FORS spectroscopic data, the model
intensities at each wavelength were scaled to the actual
counts in the utilized spectra for GJ 1214 to account for
the unique bandpass and flux weighting of the resulting
photometry. For the HAWKI photometric data, we es-
timated the effective observed stellar spectrum by mul-
tiplying the baseline theoretical stellar spectrum by a
theoretical model for the transmission of Earth's atmo-
sphere and by the transmission curve of the utilized fil-
ter. The theoretical model for the transmission of Earth's
atmosphere was calculated for the line of sight through
atmosphere above the observatory on the night of the ob-
servations using the method described by Seifahrt et al.
(2010). After integrating over the effective bandpasses,
we fit a quadratic function to the intensities to determine
the limb darkening coefficients. We limited our fits to the
range µ ≡ cos θ < 0.1, where θ is the angle between the
emergent intensity and the line of sight.
We allowed the limb darkening coefficients to be free
parameters in the light curve modeling, and we used
the estimated theoretical values as priors. That is, the
differences between the coefficients used for the light-
curve modeling and the coefficients calculated from the
8
Bean et al.
Fig. 8. -- Left panel Normalized light curves from the FORS blue data (circles) and best-fit models (lines). Right panel Residuals from
the fits (circles).
PHOENIX models were included in the tabulation of the
goodness-of-fit metric (see below). The adopted uncer-
tainties for the priors were set by the spread in the values
for the coefficients at the three different temperatures
considered for GJ 1214. We also tested the effects of
using the Claret (2000) non-linear limb darkening law.
While this functional form yielded better fits at µ < 0.1
for the MMIRS data, it did not yield better light curve
fits and the final determined transmision spectrum val-
ues for the planet are not significantly affected by using it
instead of the quadratic law. The best-fit limb darkening
coefficients were in all cases within 2 σ of the theoretical
values estimated from the Tef f = 3026 K model atmo-
sphere.
We first analyzed the broadband light curves (MMIRS
J, H, and K; FORS blue and red; and HAWKI) to de-
termine values for all the transit parameters. We ana-
lyzed the MMIRS, FORS, and HAWKI data separately.
We assumed the system scale, inclination, and transit
time are the same for the J-, H-, and K-band MMIRS
data because they were obtained simultaneously for the
same transit. Each of the MMIRS bands were allowed to
have a unique transit depth. We used a non-linear least-
squares algorithm (Markwardt 2009) to identify the best-
fit parameters, and a residual permutation bootstrap al-
gorithm to asses the uncertainties on the parameters.
The standard χ2 metric was used throughout to asses
the quality of the model fits.
The determined transit times from analysis of the
TABLE 3
Photometric Bandpasses & Transit Depths
for the FORS Blue Data
Band
Wavelength (nm)
Rp/R⋆
a
broadband
channel 1
channel 2
channel 3
channel 4
channel 5
channel 6
channel 7
channel 8
channel 9
channel 10
channel 11
channel 12
610
610
630
650
670
690
710
730
750
770
790
810
830
--
--
--
--
--
--
--
--
--
--
--
--
--
850
630
650
670
690
710
730
750
770
790
810
830
850
0.1178 ± 0.0007
0.1173 ± 0.0018
0.1195 ± 0.0012
0.1167 ± 0.0011
0.1194 ± 0.0011
0.1169 ± 0.0009
0.1164 ± 0.0009
0.1182 ± 0.0007
0.1187 ± 0.0008
0.1172 ± 0.0008
0.1172 ± 0.0007
0.1183 ± 0.0006
0.1168 ± 0.0007
a From an analysis with the transit parameters a/R⋆
and i fixed to 14.97 and 88.94◦, respectively.
broadband data with all parameters free are given in
Table 2. Combining these times with the previously
reported times from D´esert et al. (2011), Carter et al.
(2011), and Berta et al. (2011), we can calculate a re-
vised period P = 1.58040481 ± 1.2E-7 d, and reference
transit
time Tc = 2454966.525123± 0.000032 BJDTDB.
The new transit times in this paper deviate by less than
1.3σ from the predictions of this ephemeris, and none of
other transit times exhibit deviations greater than 2.4 σ.
The best-fit values of the system scale and the planet's
TABLE 4
Photometric Bandpasses & Transit Depths
for the FORS Red Data
Band
Wavelength (nm)
Rp/R⋆
a
broadband
channel 1
channel 2
channel 3
channel 4
channel 5
channel 6
channel 7
channel 8
channel 9
channel 10
channel 11
channel 12
channel 13
channel 14
channel 15
channel 16
channel 17
channel 18
channel 19
channel 20
channel 21
channel 22
780
780
790
800
810
820
830
840
850
860
870
880
890
900
910
920
930
940
950
960
970
980
990
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
--
1000
790
800
810
820
830
840
850
860
870
880
890
900
910
920
930
940
950
960
970
980
990
1000
0.1168 ± 0.0006
0.1167 ± 0.0009
0.1160 ± 0.0007
0.1156 ± 0.0007
0.1176 ± 0.0008
0.1176 ± 0.0007
0.1162 ± 0.0007
0.1172 ± 0.0007
0.1151 ± 0.0008
0.1168 ± 0.0007
0.1171 ± 0.0007
0.1171 ± 0.0007
0.1159 ± 0.0007
0.1167 ± 0.0006
0.1175 ± 0.0007
0.1178 ± 0.0006
0.1165 ± 0.0009
0.1168 ± 0.0008
0.1176 ± 0.0009
0.1172 ± 0.0009
0.1161 ± 0.0010
0.1165 ± 0.0011
0.1168 ± 0.0011
a From an analysis with the transit parameters a/R⋆
and i fixed to 14.97 and 88.94◦, respectively.
orbital inclination for the broadband data are consistent
with previously determined values (e.g., Carter et al.
2011; Berta et al. 2011). We fixed these parameters to
the values used by Bean et al. (2010, a/R⋆ = 14.9749,
i = 88.94◦) to measure the transit depth as a function
of wavelength in the subsequent analyses. These values
were also adopted by D´esert et al. (2011) and Croll et al.
(2011a) in their studies, so we can directly compare our
results with theirs. We also fixed all the transit times
to the predicted values (not best-fit) from the ephemeris
given above. This is justified because there is no evidence
of transit timing variations for this system.
We performed two types of analyses to measure the
planet's transmission spectrum. We used non-linear
least-squares fitting with a residual permutation boot-
strap to analyze the broadband light curves. For the
spectroscopic data sets, we used a Markov Chain Monte
Carlo (MCMC) algorithm to find the most likely param-
eter values and their uncertainties. The adopted photo-
metric errors for the MCMC analysis were the photon-
limited errors adjusted upwards to yield a reduced χ2 = 1
for the best fit. The free parameters for each wavelength
bin in the two analyses were Rp/R⋆, quadratic limb-
darkening coefficients (with priors), flux normalization,
and decorrelation parameters. There were a total of six
free parameters for each wavelength bin, and two "obser-
vations" per bin in addition to the light curves (the limb
darkening priors).
The motivation for the different analyses between the
broadband and spectroscopic data sets is that they are
used in different ways to investigate the planet's trans-
mission spectrum. The broadband data are useful when
compared to theoretical models in aggregate because
they span a wide range of wavelengths. For this kind
of study, the correlated, or red, noise in each data set
must be considered because the observations were ob-
tained for different transits with different instruments
9
Fig. 9. -- Upper panel The normalized HAWKI light curve (cir-
cles) and best-fit model (line). Lower panel Residuals from the
best-fit model (circles).
over the course of years. On the other hand, the spectro-
scopic data sets each tightly constrain the planet's trans-
mission spectrum on their own. These measurements
were obtained simultaneously with the same instrument
for the same transit and over a limited wavelength range.
In this case, the major sources of correlated noise, vari-
ations in Earth's atmospheric transparency and stellar
spot crossing, are quite similar from channel to channel
within a data set. Therefore, a full correlated noise anal-
ysis is not appropriate when the resulting planet trans-
mission spectra for each spectroscopic data set are only
examined in isolation, as we do in §4.
The determined Rp/R⋆ values for the MMIRS, FORS
blue, and FORS red data are given in Tables 1, 3, and
4, respectively. The determined Rp/R⋆ for the HAWKI
data is 0.1179 ± 0.0012. The normalized light curves with
systematics removed along with the best-fit transit mod-
els and residuals from the models are shown in Figures 7
(MMIRS data), 8 (FORS blue data), and 9 (HAWKI
data). The re-analyzed FORS red data do not look sub-
stantially different than those data shown by Bean et al.
(2010, spectroscopic data) and Berta et al. (2011, broad-
band data).
There is some indication of the planet crossing spots on
the surface of GJ 1214 in the FORS blue data. We per-
formed tests to investigate the influence of masking the
possible spot crossing events from the light curves. The
determined depths of the light curves changed depend-
ing on which points were masked, but the variation was
always less than the quoted uncertainties on the Rp/R⋆
values, and the relative values of the FORS blue trans-
mission spectrum were unchanged. The results presented
here are from an analysis with none of the light curve
points masked.
4. DISCUSSION
We compare our newly measured transit depths to
previous results and the predictions of theoretical mod-
els to investigate the nature of GJ 1214b's atmosphere
in this section. We utilize the models presented by
Miller-Ricci & Fortney (2010) throughout. The goals of
our investigation are to determine the composition of the
planet's atmosphere, and to provide an independent as-
10
Bean et al.
Fig. 10. -- Broadband measurements of the transmission spectrum of GJ 1214b (filled points) compared to theoretical models (lines). The
models were binned over the bandpasses of the measurements (open circles) and scaled to give the best fit to the data. The WIRCAM J-
and K-band points from (Croll et al. 2011a, grey asterisks) were not included in the fitting. All calculations were done with high-resolution
models; the models shown are smoothed for clarity.
sessment of the transit depth in the K-band to test the
claim from Croll et al. (2011a) that the transit is signif-
icantly deeper at these wavelengths.
We look at the broadband, and the optical and K-
band spectroscopic data sets separately. The spectro-
scopic data enable a search for spectral features within
the different regions they cover. The strength of these
data sets is that they yield extremely precise measure-
ments on an internal scale, and each one tightly con-
strains the planet's transmission spectrum. However,
there are offsets between the different spectroscopic data
sets with sizes that are on the order of the uncertain-
ties of the respective broadband points. The FORS blue
and red data can be combined because the offset between
the data sets can be determined from the values at the
wavelengths where they overlap. The FORS and MMIRS
spectroscopy can not be combined because the offset be-
tween the data sets is uncertain.
The broadband data cover a wide span of wavelengths,
and thus probe different sources of opacity and dif-
ferent atmospheric pressures. These data were taken
at different epochs and with different instruments, but
can be combined because we considered the correlated
noise in each of the data sets, and because the activity-
related photometric variability of GJ 1214 itself should
only cause variations in the transit depth that are much
smaller than the measurement uncertainties (Berta et al.
2011). The excellent agreement between the FORS blue
and red observations, which were separated by a year in
time, at the wavelengths the data sets have in common
(see §4.3) is additional evidence that stellar activity is
not a significant problem for this investigation.
4.1. Broadband photometry
The broadband measurements of GJ 1214b's transmis-
sion spectrum from D´esert et al. (2011, Spitzer data),
Croll et al. (2011a, ground-based near-infrared), and this
paper are shown compared to theoretical models for the
planet's atmosphere in Figure 10. The broadband trans-
mission spectrum from the combination of the FORS
blue and red, MMIRS, HAWKI, and Spitzer data does
not exhibit any significant features. The maximum devi-
ation from the weighted mean of the points is 1.7 σ.
Up until now, the only candidate detection of spectral
features in GJ 1214b's transmission spectrum has been in
the K-band by Croll et al. (2011a). Our HAWKI data
point in a narrow window of the K-band does not have
the precision to provide a strong test of the Croll et al.
(2011a) measurement on its own. However, the broad
K-band data point created from the combination of the
MMIRS spectroscopic data has excellent precision and
can provide this test. The MMIRS K-band point is 4.1 σ
(∆ Rp/R⋆ = 0.0041 ± 0.0010) lower than the Croll et al.
(2011a) K-band point, and is consistent with all of our
group's measurements at other wavelengths. Therefore,
we do not confirm the detection of spectral features by
Croll et al. (2011a).
The theoretical models for the planet's atmosphere
we compare to the observations were scaled to give the
best-fit to the FORS blue and red, MMIRS, HAWKI,
and Spitzer points. There are eight data points and
one free parameter, and thus seven degrees of freedom
for the model fits. A model for the planet's atmo-
sphere with a 100% water composition, which is essen-
tially flat at the level of the precision of the measure-
ments, yields χ2 = 10.1. On the other hand, a model
for the planet's atmosphere with a solar composition
(i.e., hydrogen-dominated) and assuming chemical equi-
librium gives χ2 = 83.9. This suggests that the simple
solar composition model is ruled out at 7.9 σ confidence.
This strengthens the conclusions we reached previously
using subsets of the current data set (Bean et al. 2010;
D´esert et al. 2011).
Methane is an important opacity source for solar com-
position models of GJ 1214b's cool atmosphere, partic-
ularly in the K-band and at 3.6 µm. However, the
predicted abundance of this molecule is subject to sig-
nificant uncertainty because it could be affected by
non-equilibrium processes (e.g., photochemistry, ther-
mal chemistry, and mixing). Miller-Ricci Kempton et al.
11
the data (χ2 = 0.9), while the solar composition model in
equilibrium is ruled out at 4.3 σ confidence using these
data alone (χ2 = 24.5). The main opacity source in this
window is methane, and the solar composition model
with methane artificially removed is consistent with the
data (χ2 = 2.0). The conclusion we draw from these data
is that the planet's atmosphere must either be metal-
enhanced or methane-depleted.
4.3. Optical spectroscopy
The optical spectroscopic measurements from the new
FORS blue data and the re-analyzed FORS red data are
shown with theoretical models in Figure 12. We examine
these data in isolation, and adjust the models to give the
best fit without consideration for the measurements at
other wavelengths. We find that the FORS blue values
are on average 0.0007 larger (indicating a deeper transit)
than the FORS red data for the wavelength range the
two data sets have in common. This is less than the
1 σ uncertainty (0.0011) between the broadband points
for each data set. We subtracted this determined offset
from the FORS blue data for display in Figure 12 and
the calculation of the fit quality for the different models.
The FORS optical data include 34 measurements, and
there are 33 degrees of freedom for the examination of
the model fit quality.
The maximum deviation from the weighted mean of
the FORS spectroscopic points is 2.3 σ, which suggests
that no spectral features are detected in these data.
The solar composition model
in equilibrium is ruled
out at 6.6 σ confidence (χ2 = 115.7) using these data
alone. The solar composition model with methane re-
moved is discrepant from the FORS data at the 5.2 σ
level (χ2 = 91.2). The 100% water composition model is
consistent with the data (χ2 = 32.4). Water mass frac-
tions of 70% are needed to bring the models within 3 σ
of the data, suggesting a highly metal-rich atmosphere.
This conclusion is the same as we found before using
only the FORS red data at lower resolution (Bean et al.
2010). The higher resolution and precision for the FORS
red data enabled by our improved data reduction, and
the addition of the blue data tightens the constraints on
the planet's optical transmission spectrum.
5. CONCLUSIONS
We have obtained new ground-based measurements of
the transmission spectrum of the 6.5 M⊕ planet GJ 1214b
in the optical between 0.61 and 0.85 µm, and in the J, H,
and K near-infrared windows. We have also re-analyzed
our previously reported red optical spectroscopy for the
planet between 0.78 and 1.00 µm. We were able to
push to higher resolution transmission spectrum mea-
surements with these data because of improvements in
the data reduction algorithm.
We combined the new data with previously reported
measurements for GJ 1214b's transmission spectrum that
were obtained with Spitzer (D´esert et al. 2011). The
combined data set spans the visible to the infrared
(0.6 to 4.5 µm), which makes it one of the most com-
plete exoplanet transmission spectra obtained to date.
We compared the combined data set to the previously
reported measurements of Croll et al. (2011a) and to
theoretical models for the planet's atmosphere from
Miller-Ricci & Fortney (2010). Our main conclusion is
Fig. 11. -- The derived transmission spectrum of GJ 1214b from
the MMIRS K-band spectroscopy (filled triangles) compared to
theoretical models (lines). The models were binned over the band-
passes of the measurements (open circles) and scaled to give the
best fit to the data. All calculations were done with high-resolution
models; the models shown are smoothed for clarity.
(2011)
studied this possibility and concluded that
methane depletion can at most only reach atmospheric
pressure levels of 10−4 bar, which is far above the al-
titude probed by transmission spectroscopy (approxi-
mately 0.01 to 1 bar). Similar conclusions have been
reached about the difficulty of photochemistry to sig-
nificantly alter the methane abundance at observable
depths in the atmosphere of the warmer planet GJ 436b
(Line et al. 2011). Comparing a solar-composition model
with methane artificially removed to our group's broad-
band data for GJ 1214b gives χ2 = 43.8, which is dis-
crepant at 5.2 σ.
A grey broadband optical opacity source is needed for
hydrogen-dominated models of the planet's atmosphere
to be consistent with the data. Such a source could be
high-altitude clouds or haze. The data constrain the al-
titude of clouds or haze in GJ 1214b's atmosphere to be
at least above a level corresponding to pressures of ap-
proximately 0.1 bar.
If methane is depleted down to a
similar level, then the clouds or haze would only have to
be optically thick for wavelengths less than 1 µm.
4.2. K-band spectroscopy
In addition to providing broadband measurements, the
MMIRS data also yield a low-resolution spectrum of
GJ 1214b in the K-band. These data are shown com-
pared to theoretical models in Figure 11. We examine
these data in isolation, and adjust the models to give the
best fit without consideration for the measurements at
other wavelengths. There are three degrees of freedom
for the model comparisons.
The MMIRS K-band spectrum is featureless and flat
like the broadband data. The maximum deviation from
the weighted mean of the points is 0.5 σ. So not only
do we not confirm in our broadband data the deeper K-
band transit suggested by Croll et al. (2011a), we also
do not detect any spectral features in the planet's at-
mosphere at these wavelengths using differential spec-
troscopy. Crossfield et al. (2011) previously presented a
non-detection of spectral features in this region based on
high-resolution differential spectroscopy, and we confirm
their results at lower resolution.
The 100% water atmosphere model is consistent with
12
Bean et al.
Fig. 12. -- The derived transmission spectrum of GJ 1214b from the FORS blue data (filled stars) and FORS red data (filled circles)
spectroscopy compared to theoretical models (lines). The FORS blue were adjusted downward by 0.0007 to match the red data in the
region where the data sets overlap. The models were binned over the bandpasses of the measurements (open circles) and scaled to give the
best fit to the data. All calculations were done with high-resolution models; the models shown are smoothed for clarity.
that there is no evidence of features in the planet's trans-
mission spectrum, and we do not confirm the detec-
tion of a significantly deeper transit in the K-band by
Croll et al. (2011a).
Our current knowledge of GJ 1214b's mass and ra-
dius indicates that the planet must have an atmosphere
(Rogers & Seager 2010b; Nettelmann et al. 2011). Our
interpretation of the featureless transmission spectrum
for the planet is that its atmosphere must either have at
least 70% H2O by mass or optically thick high-altitude
clouds or haze. A simple cloud- and haze-free model for
the planet's atmosphere with solar composition gas in
chemical equilibrium is ruled out at high confidence.
Alternatively, our knowledge of the planet's radius
could be wrong. If the planet were actually significantly
smaller (approximately 2 R⊕, or 6.5 σ from the current
best estimate), then it would not necessarily need to have
a substantial atmosphere. The estimate of GJ 1214b's
radius depends on the assumed distance to the system,
which is used to estimate the mass of GJ 1214 itself, and
this is the link in the chain with the biggest question
mark above it because the most recent estimate is based
on photographic plate measurements (van Altena et al.
1995). If the parallax were approximately 131 mas (10 σ
greater than the current estimate), the radius of the
planet determined from the light curves would be con-
sistent with the planet having no atmosphere. A new
measurement of the system's trigonometric parallax us-
ing modern technology would further our knowledge of
this important planet.
We thank Bryce Croll and Bjorn Benneke for discus-
sions pertaining to this work. We thank Brian McCleod
and Warren Brown for sharing their knowledge about the
MMIRS instrument, and Brice-Olivier Demory for infor-
mation about the HAWKI instrument. J.L.B. and E.K.
acknowledge funding from NASA through the Sagan Fel-
lowship Program. B.S. acknowledges supoort from NSF-
MRI grant 0723073. A.S. acknowledges support from
NSF grant AST-1108860. The results presented are
based on observations made with ESO Telescopes at the
Paranal Observatories under program 087.C-0505; and
on observations made with the 6.5 m Magellan telescopes
located at Las Campanas Observatory.
Facilities: VLT:Antu (FORS), VLT:Yepun (HAWKI),
Magellan:Clay (MMIRS)
REFERENCES
Adams, E. R., Seager, S., & Elkins-Tanton, L. 2008, ApJ, 673,
Croll, B., Jayawardhana, R., Fortney, J. J., Lafreni`ere, D., &
1160
Anderson, D. R., et al. 2010, A&A, 513, L3+
Appenzeller, I., et al. 1998, The Messenger, 94, 1
Bean, J. L., Miller-Ricci Kempton, E., & Homeier, D. 2010,
Nature, 468, 669
Berta, Z. K., Charbonneau, D., Bean, J., Irwin, J., Burke, C. J.,
D´esert, J.-M., Nutzman, P., & Falco, E. E. 2011, ApJ, 736, 12
Borucki, W. J., et al. 2011, ApJ, 736, 19
Carter, J. A., Winn, J. N., Holman, M. J., Fabrycky, D., Berta,
Z. K., Burke, C. J., & Nutzman, P. 2011, ApJ, 730, 82
Charbonneau, D., et al. 2009, Nature, 462, 891
Claret, A. 2000, A&A, 363, 1081
Croll, B., Albert, L., Jayawardhana, R., Miller-Ricci Kempton, E.,
Fortney, J. J., Murray, N., & Neilson, H. 2011a, ApJ, 736, 78
Croll, B., Albert, L., Lafreniere, D., Jayawardhana, R., &
Fortney, J. J. 2010a, ApJ, 717, 1084
Albert, L. 2010b, ApJ, 718, 920
Croll, B., Lafreniere, D., Albert, L., Jayawardhana, R., Fortney,
J. J., & Murray, N. 2011b, AJ, 141, 30
Crossfield, I. J. M., Barman, T., & Hansen, B. M. S. 2011, ApJ,
736, 132
D´esert, J.-M., et al. 2011, ApJ, 731, L40+
Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935
Gibson, N. P., et al. 2010, MNRAS, 404, L114
Gillon, M., et al. 2009, A&A, 506, 359
Hauschildt, P. H., Allard, F., Ferguson, J., Baron, E., &
Alexander, D. R. 1999, ApJ, 525, 871
Howard, A. W., et al. 2011, ApJ in press, arXiv:1103.2541
Kissler-Patig, M., et al. 2008, A&A, 491, 941
Line, M. R., Vasisht, G., Chen, P., Angerhausen, D., & Yung,
Y. L. 2011, ApJ, 738, 32
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Markwardt, C. B. 2009, in Astronomical Society of the Pacific
Moehler, S., Freudling, W., Møller, P., Patat, F., Rupprecht, G.,
Conference Series, Vol. 411, Astronomical Data Analysis
Software and Systems XVIII, ed. D. A. Bohlender, D. Durand,
& P. Dowler, 251 -- +
McLeod, B. A., Fabricant, D., Geary, J., Martini, P., Nystrom,
G., Elston, R., Eikenberry, S. S., & Epps, H. 2004, in Society of
Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, Vol. 5492, Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference Series, ed. A. F. M. Moorwood &
M. Iye, 1306 -- 1313
Miller-Ricci, E., & Fortney, J. J. 2010, ApJ, 716, L74
Miller-Ricci, E., Seager, S., & Sasselov, D. 2009, ApJ, 690, 1056
Miller-Ricci Kempton, E., Zahnle, K., & Fortney, J. J. 2011, ApJ
submitted, arXiv:1104.5477
& O'Brien, K. 2010, PASP, 122, 93
Nettelmann, N., Fortney, J. J., Kramm, U., & Redmer, R. 2011,
ApJ, 733, 2
Rogers, L. A., & Seager, S. 2010a, ApJ, 712, 974
-- . 2010b, ApJ, 716, 1208
Sada, P. V., et al. 2010, ApJ, 720, L215
Seifahrt, A., Kaufl, H. U., Zangl, G., Bean, J. L., Richter, M. J.,
& Siebenmorgen, R. 2010, A&A, 524, A11+
van Altena, W. F., Lee, J. T., & Hoffleit, E. D. 1995, The general
catalogue of trigonometric [stellar] parallaxes, ed. van Altena,
W. F., Lee, J. T., & Hoffleit, E. D.
13
|
1903.12116 | 1 | 1903 | 2019-03-28T17:01:33 | An intense thermospheric jet on Titan | [
"astro-ph.EP"
] | Winds in Titan's lower and middle atmosphere have been determined by a variety of techniques, including direct measurements from the Huygens Probe over 0-150 km, Doppler shifts of molecular spectral lines in the optical, thermal infrared and mm ranges, probing altogether the ~100-450 km altitude range, and inferences from thermal field over 10 mbar - 10 -3 mbar (i.e. ~100-500 km) and from central flashes in stellar occultation curves. These measurements predominantly indicated strong prograde winds, reaching maximum speeds of ~150-200 m/s in the upper stratosphere, with important latitudinal and seasonal variations. However, these observations provided incomplete atmospheric sounding; in particular, the wind regime in Titan's upper mesosphere and thermosphere (500- 1200 km) has remained unconstrained so far. Here we report direct wind measurements based on Doppler shifts of six molecular species observed with ALMA. We show that unlike expectations, strong prograde winds extend up to the thermosphere, with the circulation progressively turning into an equatorial jet regime as altitude increases, reaching ~340 m/s at 1000 km. We suggest that these winds may represent the dynamical response of forcing by waves launched at upper stratospheric/mesospheric levels and/or magnetospheric-ionospheric interaction. We also demonstrate that the HNC distribution is restricted to Titan's thermosphere above ~870 km altitude. | astro-ph.EP | astro-ph | An intense thermospheric jet on Titan
E. Lellouch1, M.A. Gurwell2, R. Moreno1, S. Vinatier1, D.F. Strobel3,
A. Moullet4, B. Butler5, L. Lara6, T. Hidayat7, E. Villard8
1. LESIA, Observatoire de Paris, Université PSL, CNRS, Sorbonne Université, Univ. Paris Diderot, Sorbonne
Paris Cité, 5 place Jules Janssen, 92195 Meudon, France
2. Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA
3. Departments of Earth & Planetary Sciences and Physics & Astronomy, Johns Hopkins University, 3400
N. Charles Street, Baltimore, MD 21218, USA
4. SOFIA / USRA, NASA Ames Research Center, Moffett Field, CA 94035, USA
5. National Radio Astronomical Observatory, Socorro, NM 87801, USA
6.
7.
8.
Instituto de Astrofisica de Andalucia -- CSIC, Glorieta de la Astronomia, E-18008 Granada, Spain
Bandung Institute of Technology, Bandung, West Java, Indonesia
Joint ALMA Observatory, Alonso de Cordova, 3107 Vitacura, Santiago, Chile
Winds in Titan's lower and middle atmosphere have been determined by a variety
of techniques, including direct measurements from the Huygens Probe1 over 0-150
km, Doppler shifts of molecular spectral lines in the optical, thermal infrared and
mm ranges2-4, probing altogether the ~100-450 km altitude range, and inferences
from thermal field over 10 mbar - 10-3 mbar (i.e. ~100-500 km)5-6 and from central
flashes in stellar occultation curves7-9. These measurements predominantly
indicated strong prograde winds, reaching maximum speeds of ~150-200 m/s in
the upper stratosphere, with important latitudinal and seasonal variations.
However, these observations provided incomplete atmospheric sounding; in
particular, the wind regime in Titan's upper mesosphere and thermosphere (500-
1200 km) has remained unconstrained so far. Here we report direct wind
measurements based on Doppler shifts of six molecular species observed with
ALMA. We show that unlike expectations, strong prograde winds extend up to the
thermosphere, with the circulation progressively turning into an equatorial jet
regime as altitude increases, reaching ~340 m/s at 1000 km. We suggest that these
winds may represent the dynamical response of forcing by waves launched at
upper stratospheric/mesospheric levels and/or magnetospheric-ionospheric
interaction. We also demonstrate that the HNC distribution is restricted to Titan's
thermosphere above ~870 km altitude.
Observations of Titan were conducted on July 22 and August 18/19, 2016
with the Atacama Large Millimeter Array (ALMA). The period corresponded to Titan's
Northern late Spring (solar longitude~ 80°). Observations altogether covered a fraction
of the 329-364 GHz range, by means of three scheduling blocks, each associated with a
frequency setup (see details in Supplementary Information). ALMA was at that time in a
moderately extended configuration (C40-5), providing angular resolutions of ~0.16-0.23"
("natural weighting"), to be compared to 1.0"-1.05" for Titan and its ~1000 km thick
atmosphere. The spectra, taken at spectral resolutions of 1-2 MHz, include the rotational
signatures of nine molecular species (CO, HCN, HNC, HC3N, CH3CN, C2H5CN, C2H3CN,
CH3CCH and C3H8, some of them with several isotopologues), mostly present in emission
due to their being formed in Titan's stratosphere. As expected based on radiative-transfer
model calculations, spectral signatures (see examples in Supplementary Figure 2) are
1
best apparent at the limb, in relation with the gain in path and the decrease of the N2
continuum. In addition, as already seen in previous ALMA and other datasets4,10-11, most
molecular emissions show non-uniform distributions across the disk. This results from
the combined effects of latitudinal variations of temperature5-6 and composition and of
the observing geometry. In this manner our data bear information of the 3D temperature
and composition fields, complementing the Cassini results12-13 for some of the molecules
(CO, HCN, HC3N, CH3CCH) and providing new information for the others.
Six molecules (HCN, DCN, HNC, HC3N, CH3CN, and CH3CCH) were further observed at a
high spectral resolution of ~60 kHz (~280 kHz for CH3CCH) in order to map their precise
line-of-sight (LOS) central frequencies. As exemplified on Fig. 1 for HNC and CH3CN,
strong blue (red) shifts are evident at beam positions centered on Titan's east (west)
equatorial limbs at 425 km altitude. These shifts correspond to a beam-integrated limb-
to-limb speed difference of ~490 m/s for HNC and ~400 m/s for CH3CN. This is much
larger than Titan's solid rotation (16 m/s at 1000 km altitude) and implies very fast
prograde winds. Maps of these Doppler winds are shown in Fig. 2 for all six species
(including separate fits for the components of the HCN hyperfine triplet), based on
Gaussian fits of the line central frequencies at all beam positions with sufficient line signal
(see details in Supplementary Information). For most molecules, wind information is
restricted to regions at or beyond the Titan limb, as emission features are too weak in
nadir geometry. A prominent exception is CH3CN, for which the strong extended emission
and the availability of three lines enable LOS wind measurements over the entire disk
with an accuracy of ~6-8 m/s. For other molecules, LOS wind errors (at/beyond limb)
are 6-8 m/s for HC3N and HCN, 12-20 m/s for HNC, and 30-50 m/s for the CH3CCH and
DCN that exhibit weaker and broader lines.
Very strong prograde winds are seen in all species, with maximum measured LOS winds
of 250-350 m/s, and even the less precise CH3CCH and DCN winds confirm this picture.
Fig. 2 reveals however differences in the wind regime from the different species. HNC LOS
speeds are strongest at equator, forming a jet, with no hint for latitudinal asymmetry. In
contrast, CH3CN winds are enhanced in the Southern (winter) hemisphere, as seen by the
"widening" of the iso-wind contours when moving from South to North. None of the
studied molecules shows convincing evidence for meridional or day-to-night flows, with
upper limits estimated to be 25-30 m/s.
Assigning altitudes/pressure levels to these wind measurements relies on the so-called
contribution functions, themselves dependent on the local vertical distributions of
temperature and abundances and the pointing geometry. Ultimately, this would require
analyzing our entire dataset in terms of the atmospheric thermal/composition 3D field.
For now, we calculated typical contribution functions for equatorial nadir and limb
conditions. Details on the procedure, involving inversion techniques for temperatures
and abundances, are given in Methods and Supplementary Information. Note that for
species whose lines are mostly Doppler-broadened (HNC and HC3N), the classical
inversion technique based on line profile was complemented by the limb sounding
capabilities offered by the data. This novel approach provides us with the first
determination of the HNC profile in Titan's atmosphere, found to be essentially restricted
to thermospheric levels above ~870 km altitude. This has important implications for
models14; in particular, the lack of HNC in the stratosphere suggests that Galactic Cosmic
Rays play at most a minor role for Titan's HNC chemistry.
2
Equatorial profiles for all six species and the associated contribution functions are shown
in Fig. 3. Together the various molecules probe the entire atmosphere above 150 km,
encompassing the mesosphere (300-800 km) and thermosphere (> 800 km). The latter is
best sounded by HNC which probes 900-1100 km, providing the first direct zonal wind
measurement in that part of Titan's atmosphere. CH3CN, DCN and CH3CCH sound the
upper stratosphere/lower mesosphere at 150-400 km, while, in relation with their steep
slope and large thermospheric abundance, HCN (main line at 354.505 GHz) and HC3N
sound a broad region roughly centered at ~700 km and extending 200-300 km above and
below, as detailed in Methods. The difference in the sounded altitudes is consistent with
the similarities/differences in the structure of the winds for the different molecules.
Supplementary Figure 6 shows LOS winds as a function of projected distance from
Titan's rotation axis. In spite of their larger uncertainties, winds inferred from DCN and
CH3CCH have a similar structure as those from CH3CN. In contrast, HNC data show a
distinct signature of high speed winds at large projected distances from the polar axis,
corresponding to the equatorial jet outlined above. Winds derived from HC3N and HCN
bear similar structures, that appear intermediate between the previous ones (i.e. stronger
winds than those from CH3CN, but with a less prevalent jet compared to HNC). Simple fits
of the LOS winds that assume solid-body rotation indicate that winds increase with
altitude, from ~200 m/s equatorial at ~350 km to ~280-300 m/s at 600 km and above.
The Richardson number associated with this wind profile remains comfortably above the
Ri < ¼ limit for turbulence.
Assuming purely zonal flow, we used our forward model and a trial-and-error approach
to de-convolve the LOS winds measured in CH3CN, HNC and HC3N in terms of latitudinal
wind profiles at the associated altitudes (see Supplementary Information). Results for
CH3CN, which pertain to 345+60-140 km (0.01-0.7 mbar with peak sensitivity at 0.035 mbar)
indicate a maximum 220 m/s speed at 15°S with, as expected from qualitative data
inspection, a much more gradual decline in the Southern vs. Northern hemisphere. This
result agrees within error bars with pioneering measurements4 using the same species
(160±60 m/s equatorial at 300±100 km, assuming solid-body rotation). The north/south
asymmetry that is evident in our CH3CN data also confirms the well-trusted but still
indirect inferences based on temperature fields and the use of the thermal wind equation.
Those consistently indicate a strong, broad mid-latitude (~20°-50°) jet in the winter
hemisphere with peak velocities in excess of 190 m/s, extending over ~1 decade in
pressure centered at 0.1 mbar, while significant winds (up to 100 m/s) are still present at
mid-latitudes of the opposite hemisphere5-6. Note that the thermal wind approach is valid
only outside an altitude-dependent (but typically ± [10°-30°]) equatorial band, a
limitation that does not hold in direct wind measurements. Analyses of stellar occultation
central flashes in 2001-2003 (Northern winter solstice) also indicate maximum winds of
~200 m/s at 250 km, but disagree on their latitudinal extent7-9.
For HC3N and HNC, the de-convolved equatorial speeds are 310-345 m/s, rapidly
decreasing with latitude in both hemispheres and with no clear latitudinal asymmetry.
For HC3N, results are sharply different from Ref. [4], both in the estimate in the probed
altitude (710+190-260 km here vs 450±100 km in [4]) and in the wind velocities (310 m/s
equatorial here, vs 60±20 m/s in [4]). As the data from [4] refer to 2003-2004 (solar
longitude L =268°-280°), the contradiction -- if taken at face value -- would imply a
temporal variation of the mesospheric circulation, with winds decreasing (resp.
3
increasing) with altitude at southern (resp. northern) summer solstice, or near perihelion
(resp. aphelion).
Based on INMS and CAPS/IBS data, ion winds at 1000-1400 km in the range 30-260 m/s
(vector component along the spacecraft trajectory) were reported15. Above 1200 km, ions
and neutrals are collisionally-coupled16, so these speeds represent neutral winds, but the
measured range was broad and the other wind components unconstrained. Our HNC data
reveal that the circulation regime over 900-1100 km is characterized by a relatively broad
(±30-35° latitude extent at half maximum) equatorial jet with ~350 m/s maximum speed,
i.e. ~1.4 times the speed of sound. Such a regime was previously unexpected. Pre-Cassini
models17-18 predicted subsolar to anti-solar (SSAS) winds of ~50 m/s. Early Cassini/INMS
density profiles were used19 to construct an empirical altitude-latitude temperature
model; solving the momentum equations, a circulation pattern with a zonal jet of ~50 m/s
and very strong meridional winds (~150 m/s) at 70°N latitude was inferred. A more
complete analysis of 32 Cassini/INMS passes20 indicated that while mean temperatures
over 1050-1500 km vary over 112-175 K, there is a conspicuous lack of correlation
between temperatures, latitude, longitude, local time, and solar zenith angle, indicating
that the thermospheric thermal structure and dynamics are not primarily controlled by
the absorption of solar EUV. Some correlations were found between atmospheric
temperatures and the local plasma environment21,20, but appear to be of ambiguous
interpretation. An analysis of thermospheric energetics22 demonstrated that particle
sources (magnetospheric ions, pick-up ions, and auroral electron precipitation) and Joule
heating contribute lower power than EUV input -- typically by a factor 10 below 1100 km,
and solar EUV also dominates ionospheric production23. But Saturn's co-rotation electric
field that drives magnetospheric plasma convection past Titan at 120 km/s also likely
drives ~1 km/s ion convection in the ionosphere24. Collisionally-coupled ion/neutral
winds of ~350 m/s would appear sufficient to power perpendicular electric fields E⟂ = -
v x B of order 1.5 µV/m, as inferred25.
Vertically propagating waves seem capable of depositing energy fluxes comparable to
solar UV [Ref. 22]. Evidence for waves in Titan's mesosphere and thermosphere is
abundantly present in stellar occultation profiles9, in the Huygens/HASI profile26 above
500 km, and in the INMS density/temperature profiles22,27, and maybe in vertical
molecular profiles. Gravitational tides28 are not promising for explaining the strong wind
speeds we measure, due to their small phase velocities (0 and ± 5.9 m/s equatorial) and
inability to penetrate regions where their phase speed equals the zonal velocity (i.e.
between 0-5 km and the 70-80 km region of minimum wind1). Above 300 km altitude
where the radiative time constant becomes shorter than a Titan day29 (= 16 Earth days),
other gravity waves, in particular thermal tides generated by solar heating, diurnal and
especially semi-diurnal with large vertical wavelength30,31, will propagate vertically. They
can deposit eddy momentum fluxes by viscous dissipation, interact with the weak solar-
induced SSAS thermospheric circulation and produce a strong equatorial flow. Solar
forcing of tides will be complex as a solar day for a parcel advecting with zonal winds is
16 days at the surface, approximately one day at 350 km and ~ 3/4 day at 1000 km.
Detailed models are needed to assess the plausibility of the mechanism, but an analogy
may hold with Venus' upper atmosphere, where gravity waves are thought to be
responsible for the variable retrograde super-rotating zonal flow that superimposes to
the strong SSAS regime32.
4
References
1. Folkner, W.M., et al. Winds on Titan from ground-based tracking of the Huygens probe. J. Geophys. Res.
111, E07S02 (2006).
2. Luz, D., et al. Characterization of the zonal winds in Titan's stratosphere of Titan with UVES: 2.
Observations coordinated with the Huygens Probe entry. J. Geophys. Res. 111, E08S90 (2006).
3. Kostiuk, T. et al. High spectral resolution infrared studies of Titan: winds, temperature and composition.
Planet. Space Sci 58, 1715-1723 (2010).
4. Moreno, R., Marten, A. & Hidayat T. Interferometric measurements of zonal winds on Titan. Astron.
Astrophys. 437, 319-328 (2005).
5. Achterberg, R. K. et al. Titan's middle atmospheric temperatures and dynamics observed by the Cassini
Composite Infrared Spectrometer. Icarus 194, 263-277 (2008).
6. Achterberg, R. K. et al. Temporal variations of Titan's middle atmospheric temperatures from 2004 to
2009 observed by the Cassini/CIRS. Icarus 211, 686-698 (2011).
7. Hubbard, W.B. et al. The occultation of 28Sgr by Titan. Astron. Astrophys. 269, 541-563 (2005).
8. Bouchez, A.H. Seasonal trends in Titan's atmosphere: haze, winds, and clouds. Ph.D. Thesis, California
Institute of Technology (2004)
9. Sicardy, B. et al. The two Titan stellar occultations of 14 November 2003. J. Geophys. Res. 111, E11S91
(2006).
10. Cordiner, M. et al. ALMA measurements of the HNC and HC3N distributions in Titan's atmosphere.
Astrophys. J. Lett. 795, L30, 6 pp. (2014)
11. Lai, J. C.-Y. et al. Mapping vinyl cyanide and other nitriles in Titan's atmosphere using ALMA
Astronomical J. 154, L206, 10 pp. (2017).
12. Vinatier, S. et al. Vertical abundance profiles of hydrocarbons in Titan's atmosphere at 15°S and 80°N
retrieved from Cassini/CIRS spectra. Icarus 188, 120-138 (2007).
13. Teanby, N. et al. Vertical profiles of HCN, HC3N and C2H2 in Titan's atmosphere derived from
Cassini/CIRS data. Icarus 186, 364-384 (2007).
14. Dobrijevic, M., Loison, J.-C., Hickson, K. M. & Gronoff, G. 1D-coupled photochemical model of neutrals,
cations and anions in the atmosphere of Titan. Icarus 268, 313-339 (2016).
15. Crary, F. J., Magee, B. A., Mandt, K., Waite, J. H., Westlake, J., Young, D. T. Heavy ions, temperatures and
winds in Titan's ionosphere: Combined Cassini CAPS and INMS observations. Planet. Space Sci. 57,
1847-1856 (2009).
16. Cravens, T. et al. Dynamical and magnetic field time constants for Titan's ionosphere: Empirical
estimates and comparisons with Venus. J. Geophys. Res. 115, A08319 (2010).
17. Rishbeth, H., Yelle, R.V. & Mendillo, M. Dynamics of Titan's thermosphere. Planet. Space Sci 48, 51-58
(2000).
18. Müller-Wodarg, I.C.F., Yelle, R. V., Mendillo, M. & Aylward, A.D. On the global distribution of neutral gases
in Titan's upper atmosphere and its effect on the thermal structure. J. Geophys. Res. 108, A12, id 1453
(2003).
19. Müller-Wodarg, I. C. F., Yelle, R. V., Cui, J. & Waite, J. H. Horizontal structures and dynamics of Titan's
thermosphere J. Geophys. Res. 113, E10, id 10005 (2008).
20. Snowden, D., Yelle, R. V., Cui, J., Wahlund, J.-E., Edberg, N. J. T., Ågren, K. The thermal structure of Titan's
upper atmosphere, I: Temperature profiles from Cassini INMS observations. Icarus 226, 552-582 (2013).
21. Westlake, J. H. et al. Titan's thermospheric response to various plasma environments. J. Geophys. Res.
116, A3, id A03318 (2011).
22. Snowden, D. & Yelle, R. V. The thermal structure of Titan's upper atmosphere, II: Energetics. Icarus 228
64-77 (2014).
23. Ågren, K., Wahlund, J.-E., Garnier, P., Modolo, R., Cui, J., Galand, M., Müller-Wodarg, I. On the ionospheric
structure of Titan. Planet. Space Sci. 57, 1821-1827 (2009).
24. Ulusen, D et al. Comparisons of Cassini flybys of the Titan magnetospheric interaction with an MHD
model: evidence for organized behavior at high altitudes. Icarus, 217, 43-54 (2012).
25. Ågren, K, et al. Detection of currents and associated electric fields in Titan's ionosphere from Cassini
data. J. Geophys. Res. 116, A04313 (2011).
26. Aboudan, A., Colombatti, G., Ferri, F. & Angrilli, F. Huygens probe entry trajectory and attitude estimated
simultaneously with Titan's atmospheric structure by Kalman filtering. Planet. Space Sci. 56, 573-585
(2008).
27. Müller-Wodarg I. C. F., Yelle, R. V., Borggren, N. & Waite, J. H. Waves and horizontal structures and
dynamics of Titan's thermosphere J. Geophys. Res. 111, A12, id A12315 (2006).
5
28. Strobel, D.F. Gravitational tidal waves in Titan's upper atmosphere. Icarus 182, 251-258 (2006).
29. Bézard, B., Vinatier, S., & Achterberg, R. Seasonal radiative transfer modeling of Titan's stratospheric
temperatures at low latitudes. Icarus 302, 437-450 (2018).
30. Fels, S.B. & Lindzen, R.S. The interaction of thermally excited gravity waves with mean flows. Geophys.
Fluid Dyn. 6, 149-191. (1974)
31. Zhu, X. Maintenance of equatorial superrotation in the atmosphere of Venus and Titan. Planet Space Sci.
54, 761-773 (2006)
32. Hoshino, N., Fujiwara, H., Takagi, M. & Kasaba, Y. Effects of gravity waves on the day-night difference of
the general circulation in the Venusian lower thermosphere. J. Geophys. Res. Planets 118, 2004-2015 (2013).
33. Thelen, A. et al. Abundance measurements of Titan's stratospheric HCN, HC3N, C3H4 and CH3CN from
ALMA observations. Icarus 319, 417-432 (2019).
Corresponding author: Emmanuel Lellouch. LESIA, Observatoire de Paris, Université
PSL, CNRS, Sorbonne Université, Univ. Paris Diderot, Sorbonne Paris Cité, 5 place Jules
Janssen, 92195 Meudon, France. [email protected]; 33-1-45077672.
Acknowledgements E.L., R.M., and SV acknowledge support from the Programme
National de Planétologie (PNP-INSU). We thank Agustín Sánchez-Lavega and an
anonymous reviewer for constructive comments. The paper is dedicated to the memory
of Daniel Lellouch, deceased on February 13, 2019.
Author Contributions M.G. wrote the ALMA proposal. M.G. and R.M. reduced the ALMA
data. E.L. analyzed and modelled the data (with contribution from R.M.) and wrote most
of the manuscript. S.V. provided initial thermal profiles as well as insight in the general
science context. D.F.S. led the interpretative part. All authors discussed the manuscript.
Materials requests & Correspondence should be addressed to the corresponding
author.
Competing interests The authors declare no competing interests.
6
Fig. 1. Evidence for zonal winds. CH3CN (left) and HNC (right) line profiles at Titan East
(+0.425", 0" from disk center) and West (-0.425",0" from disk center) limbs, observed on
August 18, 2016. These pointings correspond to a radial distance of ±3000 km from disk
center (i.e. a limb tangent altitude of 425 km above the surface). The reference zero
frequencies are the rest frequencies of the CH3CN J, K=18,1 line (349.44699 GHz) and HNC
(4-3) line (362.63030 GHz) respectively. The left panel also includes the CH3CN J, K=18,0
line at 349.45370 GHz. The limb-to-limb velocity difference is ~400 m/s for CH3CN and
~490 m/s for HNC.
7
Fig. 2. Doppler wind maps. Line-of-sight Doppler shifts (m/s) measured in CH3CN,
CH3CCH, DCN, HCN (main line of the hyperfine triplet at 354.5055 GHz and satellite lines
at 354.5039 and 354.5075 GHz), HC3N, and HNC. Right Ascensions (RA) and declinations
are shown in terms of Titan's apparent radius. Ellipses show the synthetized beams.
White areas correspond to regions where emission signals are too faint for reliable
measurements to be performed. Only CH3CN emission makes it possible to measure winds
throughout the disk. Some measurements within the disk are also possible using the main
HCN 354.5055 GHz line, that appears in absorption in nadir geometry (see Supplementary
Figure 2). Doppler shifts in the various species clearly indicate a prograde wind regime,
but with different structure for CH3CN vs HNC, HCN and HC3N. Winds in CH3CCH are
mostly measurable outside of the low-latitude regions. Winds are only marginally
detected from DCN emission, with 30-50 m/s error bars. See Supplementary Information
for details.
8
Fig. 3. Molecular abundance profiles and probed altitudes. Left: Equatorial vertical
profiles for the six species used for the wind measurements (solid lines, with error bars
indicated in the altitude ranges where information is available). As explained in
Supplementary Information, uncertainties on HNC and on the upper part of the HC3N
distribution are best described by a vertical shift of their altitude profile (dashed lines).
For HCN, HC3N and CH3CCH, characteristic profiles from Cassini/CIRS, UVIS and INMS,
and from previous 2012-2015 ALMA observations (Thelen et al. 2019, Ref. [33]) are
indicated. See Methods and Supplementary Information for details. Right: Wind
contribution functions (WCF) in limb (solid lines) and nadir (dashed lines) geometries.
For HCN, WCF are shown separately for the main line at 354.5055 GHz and for the satellite
lines at 354.5039 and 354.5075 GHz (averaging the two). For HC3N, to avoid the
contribution from the nearby HCN (4-3) line, a WCF calculated with zero HC3N has been
subtracted from the WCF before plotting. For other species, the peak below 100 km is due
to the continuum. See Methods for details.
9
Methods
1) Radiative transfer models
Spectral calculations were based on a home-made radiative transfer model in full
spherical geometry, inherited from [4, 34, 35]. Line opacities are calculated using
spectroscopic parameters from the Cologne Database for Molecular Spectroscopy (CDMS,
https://www.astro.uni-koeln.de/cdms). As a consequence of small particle size and
sharply decreasing absorption coefficients beyond 100 µm36, thermal emission and
scattering by haze are neglected, as in all previous studies of Titan at mm/submm
wavelengths. Pencil-beam radiances are calculated on a grid of impact parameters
(distance to Titan center) ranging from 0 to 4120 km (1.6 Titan radius). For each beam
position under study, line-of-sight radiances as a function of impact parameter from
Titan's center are re-interpolated over the Titan + atmosphere disk, and integrated over
portions of the disk intercepted by the ALMA synthesized beam. At this step, we took into
account the beam's elliptical shape and its orientation, and the precise apparent radius of
Titan for each individual observation. This resulted in synthetic spectra expressed in
Jy/beam which could then be directly compared to the data, after applying whenever
needed minor (< 10 %) model-based calibration corrections to the data fluxes.
2) Temperature and abundance retrievals
Temperature profile information was derived from modelling of the CO (3-2) and HCN (4-
3) lines and their isotopic variants, using a home-made inversion code35 and based on
optimal estimation methods principles37-38. For the purpose of this task, which is mainly
to assign altitude levels to the wind measurements, we focused on the equatorial region,
leaving the study of the thermal and compositional field for future work. For the a priori,
we used an equatorial profile derived from Cassini/CIRS measurements in May 2016
(T119 flyby), extended above 700 km at a 153 K temperature, typical of the upper
atmosphere according to Cassini / INMS measurements, and shown in Supplementary
Figure 1 (black line). While CO lines constrain atmospheric temperatures below ~350
km, HCN probes much higher altitudes (up to ~1000 km), with the difficulty that the
temperature profile and the HCN profile are poorly constrained above the top of the
sensitivity range of CIRS (~520 km). To alleviate the ambiguity between abundance and
temperature effects in the core of the HCN (4-3) line, we inverted simultaneously the
equatorial HCN (4-3) and H13CN (4-3) lines near the central meridian and at the limb.
Indeed, the characteristic absorption / emission appearance of the HCN hyperfine
structure in nadir / limb geometry provides complementary information on temperature
and abundance sensitivity. In doing so, we used the HCN profile from Ref. [39] (extended
above 400 km to reach HCN = 10-4 at 800 km and above) as a priori.
The best fit equatorial thermal profile retrieved in this manner is shown in
Supplementary Figure 1 (blue line), and the associated fits of the CO, HCN and H13CN
lines in nadir and limb geometry are shown in Supplementary Figure 2. The
temperature profile shows a marked mesopause at 140 K near 800 km. The occurrence of
such a cold temperature somewhere above the stratopause is in fact directly implied by
the depth of the main component of the HCN (4-3) hyperfine structure, with a brightness
10
temperature of 143 K in the line core in nadir geometry. Note also that the best fit
temperature / HCN profile permits a good match of the HCN v2=1 emission at 354.460
GHz (Supplementary Figures 2 and 3). The retrieved profile is also warmer than the a
priori (CIRS) profile by up to ~3 K at 250 km and colder by ~6 K at 400-600 km.
Differences between CIRS- and ALMA CO-derived thermal profiles of up to 5 K below 300
K have already been noted40.
Error bars on the retrieved HCN/temperature retrievals are a combination of (i)
uncertainties due to measurement noise and (ii) uncertainties due to the a priori used in
the inversion process. Noise-related uncertainties were calculated using the formalism
described in Ref. [37] (their eq. 23 and 24). The associated temperature errors are
formally very small (0.2-0.3 K) in relation to the high S/N and high spectral-resolution of
the data. Uncertainties due to the a priori were evaluated by repeating the
temperature/HCN inversion process using several a priori temperature and HCN profiles,
as detailed in Supplementary Information. These modelling variants yielded a suite of
solution profiles, whose dispersion is a good indicator on the actual error bars on the
retrieved temperature and HCN profile (Supplementary Figure 4). Beyond the well-
known characteristics of Titan's middle atmosphere thermal structure such as the
stratopause near 270 km, robust new features of all retrieved profiles are (i) the existence
of a quasi-isothermal region over 450-600 km (already apparent in the CIRS-based a
priori) and (ii) the existence of a ~140 K mesopause, located at 800±50 km altitude
depending on the a priori. Error bars on the retrieved temperatures remain below 2-3 K
up to 800 km; above this level they increase with the progressive lack of sensitivity of the
measurements to ultimately reflect the a priori errors. The HCN profile (Fig. 3) is also
very well defined (to within ~10-20 %) over 100-850 km, before error bars progressively
increase. The nominal HCN mixing profile we determine reaches a 6.3 x 10-4 mixing ratio
at 1000 km, in agreement with results from both Cassini/INMS41 and Cassini/UVIS42,
respectively above and below 950 km (see Fig. 3).
The retrieved thermal profile shows a quasi-isothermal part at 156 K over 450-600 km
and a 140 K minimum near 800 km. Although the resulting oscillatory structure may be
reminiscent of the temperature perturbations seen in the Huygens/HASI data26, its
attribution to a wave is doubtful because (1) our data sample broad regions (beam size
~1000 km), and nadir and limb data inverted simultaneously are distant by ~2500 km
(2) the resulting wave amplitude and vertical wavelength would be ~5-10 K and ~300
km, respectively smaller and larger than in the HASI data (~10-20 K and ~100 km).
Instead, the temperature profile, and in particular the 800 km temperature minimum,
may be mostly radiatively-controlled. A pioneering non-LTE model43 of Titan's upper
atmosphere predicted a 135-140 K mesopause (but at 600 km), primarily due to C2H6
cooling. Revising such a model in the light of new distributions derived from Cassini and
the present work (especially for HCN, C2H6, HC3N and CH3CN) and with updated
spectroscopic databases would shed further light on this issue.
The retrieved thermal profile was then used to derive the other molecular profiles, again
for the equatorial region. The DCN profile was taken as the retrieved HCN profile,
multiplied by a constant DCN/HCN ratio of (2.5±0.2) x 10-4 [44], i.e. assuming that the
DCN profile follows that of HCN at all altitudes. CH3CN and CH3CCH were inverted using a
priori profiles respectively based on ground-based data39 (extended to 10-5 at 1000 km
and above) and CIRS observations45. For these three species, best fits of the (west) limb
11
equatorial emission are presented in Supplementary Figure 3. Error bars on CH3CN and
CH3CCH reflect the combined effects of noise, temperature profile uncertainty and a priori
errors.
The Doppler-limited line-shape of HNC J=4-3 line (Supplementary Figure 5) contains
limited information on the HNC vertical profile, except for the fact that this precludes HNC
to be present at too deep atmospheric levels. Based on HNC J=6-5, ref. [34] found that for
constant mixing profiles above some level, the cutoff altitude cannot be lower than 300
km. Using HNC data averaged over a concentric ring 2000-3350 km from Titan center, and
testing photochemical profiles14 for HNC, we found that several of these HNC profiles
overestimate the amount of HNC in the lower stratosphere. In contrast, the model in
which the production of HNC by Galactic Cosmic Ray impact is omitted is essentially in
agreement with the observed spectrum. While this line shape-based approach does not
constrain the distribution of HNC at altitudes above the pressure-broadening levels, the
latter can be determined from the observed variation of the line-integrated intensities
with radial distance. Comparing the latter with model calculations for constant HNC
profiles above some cutoff altitudes, we found that the bulk of HNC is located above
870±40 km altitude, implying that HNC is mostly a thermospheric species. The
photochemical model from Ref. [14] without GCR also provides the correct radial
dependence for the HNC emission, and was adopted as nominal profile, allowing for a
vertical shift of ±40 km in its altitude scale.
The situation is similar for the HC3N J=39-38 emission at 354.697 GHz, which -- except in
the high-northern and southern latitude regions where it is strongly enhanced -- does not
show Lorentzian wings. We performed the same study as for HNC on the radial distance
variation of the HC3N emission (using the slightly stronger J=38-37 line at 345.609 GHz
observed along with CO (3-2)), restricting ourselves to low-latitude regions. Considering
piecewise profiles, this approach yielded a best fit cutoff altitude of 630±30 km
(Supplementary Figure 5), implying that HC3N too is mostly present in Titan's upper
atmosphere, albeit less so than HNC. This is not surprising given that HC3N, as measured
by CIRS12-13 has also a measurable abundance in Titan's stratosphere and lower
mesosphere, with mixing ratios increasing from 6x10-11 -- 1x10-7 over 150-450 km.
Initializing the HC3N profile at these values and a large abundance at z > 800 km, we
determined an equatorial HC3N profile fitting both the shape of the J=39-38 line and the
radial variation of the J=38-37 lines fluxes. In contrast, the nominal HC3N profile from Ref.
[14] does not fit well any of either constraint, indicating that this profile is overabundant
at lower altitudes and under-abundant in the upper atmosphere. Our HC3N solution
profile has a 1000 km abundance as high as ~5 x 10-5. This agrees rather well with the
global value reported from Cassini/INMS46 and with the Cassini/UVIS HC3N abundance
over 850-950 km [42]. We reiterate that our conclusions pertain to the equatorial HC3N
profile only.
Fig. 3 summarizes the retrieved profiles for all six molecules used to probe the winds.
Error bars are indicated in the range where information is available. For HNC and HC3N,
uncertainties are best expressed as a vertical shift of ±40 km and ±30 km in the profiles,
for reasons explained above. These profiles are compared to characteristic profiles locally
derived from Cassini/UVIS at 6°S (T41I, Ref. [42]), INMS41,46, and CIRS. For the latter, we
show HCN and CH3CCH profiles at 9°S from May 2016 (T119, Ref. [45]); for HC3N, not
clearly detected in the latter observations, we use instead a 5°N profile from Jan. 2007
12
(T23, Ref. [47]). Also shown in Fig. 3 are disk-averaged HCN and HC3N profiles derived
from 2012- 2015 ALMA observations (Ref. [33], their Fig. 10). Fig. 3 highlights the overall
agreement between these profiles, giving in particular great confidence in our HCN and
HC3N profiles in the middle/upper atmosphere. A noteworthy difference is that "local"
molecular profiles (from Cassini CIRS and UVIS) show small-scale vertical variations (see
e.g. Figs. 3- 4 of Ref. [47], and Fig. 26 of Ref. [42]), not predicted in photochemical models14
and presumably of dynamical origin; those are particularly visible in Fig. 3 for the HCN
and HC3N profiles derived from UVIS.
3) Contribution functions and wind altitudes
The retrieved profiles for all six molecules used to probe the winds are shown in Fig. 3
(left panel). Based on these profiles, "wind contribution functions" were calculated to
estimate the range of altitudes probed by the wind measurements in each molecule. For
this, we first calculated "traditional" contribution functions B(T) ∂exp(-τ) / ∂ln(p), both
monochromatically (over a spectral interval matching the one used for the wind
measurements), and for every pencil-beam line-of-sight. At each frequency, these
contributions functions were then convolved with the synthetized beam associated to
each observation. To estimate how the probed altitude may vary with the observing
geometry, these spatially-convolved -- but still monochromatic -- contribution functions
were calculated both in nadir view (i.e. pointing to Titan's disk center) and in limb view,
defined here as a position distant from Titan's center by 1.3 Titan radius (i.e. 772.5 km
above the solid planet limb). This corresponds typically to the distance where we measure
strongest winds before signals get too faint. Finally, these spatially-convolved
contribution functions were spectrally convolved, using a weight proportional to the local
spectral slope of the synthetic spectrum48,4. For each species, these wind contribution
functions (WCF) for the nadir and limb geometries are shown in the right panel of Fig. 3.
For HCN, we show separately the WCF for the main line at 354.5055 GHz and for the two
satellite lines at 354.5039 and 354.5075 GHz (averaging their two very similar WCF).
Wind contribution functions highlight three regions: (i) a narrow (900-1100 km)
thermospheric region probed by HNC (ii) another relatively narrow region, encompassing
the upper stratosphere / lower mesosphere (200-400 km) probed by CH3CN, CH3CCH,
and DCN (iii) a much broader zone, straddling from ~400 to ~1000 km (in limb
geometry), probed by HCN and HC3N. In limb geometry, the mean probed altitudes (i.e.
the WCF-weighted mean altitude) are: 345+60 -140 km for CH3CN; 315+90 -160 km for CH3CCH;
340+90 -155 km for DCN; 745+365 -320 km for the main HCN line; 540+120 -185 km for the HCN
satellite lines; 710+190 -260 km for HC3N; and 990+100 -120 km for HNC, where the "error bars"
mark the vertical extent of each WCF, i.e. the range in which it exceeds ½ of its maximum
value. The impact of the thermal profile and abundance uncertainties (see Supplementary
Information) on the wind contribution function (WCF) are small. Specifically, the mean
probed altitude, as defined above, shifts by ±45 km for HNC, ±20 km for HC3N, ±25 km
(resp. ±15 km) for the main (resp. the satellite) HCN line, and ±10 km for CH3CN, CH3CCH
and DCN. Thus, except to some degree for HNC, the associated error bar appears dwarfed
by the broad vertical extent of the WCF, from which we conclude that the vertical scale
associated with our wind measurements is robust.
13
Wind contributions functions calculated by [4] for Plateau de Bure interferometer
observations of CH3CN and HC3N lines at 220 GHz indicated sounding altitudes of
300±150 km for CH3CN and 450±100 km for HC3N. This compares well with our
computation for CH3CN but is rather different for HC3N. This stems from the combination
of three factors, all of which "pushing" the contribution function to higher levels in our
case: (i) a higher HC3N abundance in the upper atmosphere compared to the profile39
adopted in Ref. [4] (ii) the fact that we calculated the WCF for pointing towards a distance
of 1.3 Titan radius (vs 1.0 Titan radius in [4]), and (ii) the ~3 times smaller synthesized
beam in our observations (~0.2" vs 0.6"), which enhances the contribution of limb with
respect to nadir, thereby tending to probe higher altitudes.
Data availability The data that support the plots within this paper and other findings of
this study are available from the ALMA archive (http://almascience.nrao.edu/aq/) and
from the corresponding author upon reasonable request.
Code availability Novel methods developed for this research and not available from
previous studies are available from the corresponding author upon reasonable request.
Additional references
34. Moreno, R. et al. First detection of hydrogen isocyanide (HNC) in Titan's atmosphere. Astron. Astrophys.
536, id. L12, 4 pp. (2011).
35. Lellouch, E. et al. Detection of CO and HCN in Pluto's atmosphere with ALMA. Icarus 286, 289-307
(2017).
36. Anderson, C.M. & Samuelson, R.E. Titan's aerosol and stratospheric ice opacities between 18 and 500
μm: Vertical and spectral characteristics from Cassini CIRS. Icarus 212, 762-778 (2011).
37. Conrath, B.J., Gierasch, P.J., Ustinov, E.A. Thermal structure and para hydrogen fraction on the outer
planets from voyager IRIS measurements. Icarus 135, 501 -- 517 (1998).
38. Rodgers, C.D. Inverse methods for atmospheric sounding: theory and practice. Vol. 2. World Scientific,
Singapore (2000).
39. Marten, A., Hidayat, T., Biraud, Y., Moreno, R. New millimeter heterodyne observations of Titan: vertical
distributions of nitriles: HCN, HC3N, CH3CN, and the isotopic ratio 15N/14N in its atmosphere. Icarus 158,
532 -- 544 (2002).
40. Thelen, A. E. et al. Spatial variations in Titan's atmospheric temperature: ALMA and Cassini comparisons
from 2012 to 2015. Icarus 307, 380-390 (2018).
41. Cui, J., Cao, Y.-T., Lavvas, P. P., Koskinen, T. T. The variability of HCN in Titan's upper atmosphere as
implied by the Cassini Ion-Neutral Mass Spectrometer measurements. Astrophys. J. Lett. 826, L5, 6pp (2016).
42. Koskinen, T. T., Yelle, R. V., Snowden, D. S., Lavvas, P., Sandel, B. R., Capalbo, F. J., Benilan, Y., West, R. A.
The mesosphere and lower thermosphere of Titan revealed by Cassini/UVIS stellar occultations. Icarus 207,
511-534 (2011).
43. Yelle, R.V. Non-LTE models of Titan's upper atmosphere. Ap. J. 383, 380-400 (1991).
44. Molter, E.M. et al. ALMA observations of HCN and its isotopologues on Titan. Astron. J. 152, id. 42, 7 pp.
(2016).
45. Vinatier, S. et al. Seasonal variations in Titan's stratosphere observed with Cassini/CIRS during northern
spring. American Astronomical Society, DPS meeting #49, id.304.03 (2017).
46. Cui, J., et al. Analysis of Titan's neutral upper atmosphere from Cassini Ion Neutral Mass Spectrometer
measurements Icarus 200, 581-615 (2009).
47. Vinatier, S. et al. Analysis of Cassini/CIRS limb spectra of Titan acquired during the nominal mission. I.
Hydrocarbons, nitriles and CO2 vertical mixing ratio profiles. Icarus, 205, 559-570 (2010).
48. Lellouch, E., Goldstein, J.J., Rosenqvist, J., Bougher, S.W., Paubert, G. Global circulation, thermal structure,
and carbon monoxide distribution in Venus' mesosphere in 1991. Icarus 110, 315-339 (1994).
14
Supplementary Information for:
An intense thermospheric jet on Titan
E. Lellouch, M.A. Gurwell, R. Moreno, S. Vinatier, D.F. Strobel,
A. Moullet, B. Butler, L. Lara, T. Hidayat, E. Villard
correspondence to: [email protected]
1) Observations and data reduction
We observed Titan as the focus of Cycle 3 program 2015.1.01023.S (PI. M. Gurwell). Titan
was observed in three scheduling blocks (SB) named (A-C), one of them being scheduled
on July 22, 2016 and the other two consecutively on August 18/19, 2016. The quasars
J1634-2058 and J1626-2951 were used for phase calibration and flux calibration. All
three scheduling blocks made use of Band 7 receiver sets, altogether covering large
portions of the 329-364 GHz range. We employed a range of spectral resolutions.
Typically, moderate 1-2 MHz resolutions were used to capture profiles of broad lines such
as CO and HCN, while being also well suited to weak emissions. Higher resolutions (61 --
282 kHz) were also used, to study narrow yet strong lines, in particular, HNC, HC3N, and
the HCN hyperfine structure core, and with the intent of measuring winds from Doppler
shifts. The on-source time was 49 min for each of the scheduling blocks. 38 antennas were
active in July 2016 and 39 in August. Observing conditions were good (H2O = 0.44-0.66
mm) resulting in rms noise of 2.0 mJy/beam/MHz. Observing details are given in
Supplementary Table 1.
Data reduction was initially carried out under CASA. Once calibrated visibilities were
produced, deconvolution and imaging were performed under both the CASA and GILDAS
environments. Self-calibration was performed to improve imaging quality, and the data
were finally re-gridded on 0.02" or 0.05" pixel grids, expressing the spectral axis in Titan's
rest frame. The achieved angular resolution ("natural weighting") varies from almost
circular ~0.17" to more elongated 0.29" x 0.20" beams, as detailed in Supplementary
Table 1. Titan's 5150-km diameter subtended 0.7594" and 0.7283" in the sky for the July
and August observations respectively, and coordinate scales for maps were expressed in
units of Titan's radius. For both periods, the sub-earth latitude was 26.0° and the North
polar angle 3.45°.
15
Supplementary Table 1: Observational summary
Date (UT)
at start
Tint
(min)
∆ (au)1 Sub-
earth
long.
Spectral range (GHz) Spectral
resolution
(MHz)
1.13
Synthesized beam
(major,
minor,
orientation2)
0.29"x 0.20"
PA=-85°
0.29"x 0.20"
PA=-86°
0.26"x 0.19"
PA=-81°
0.28"x 0.19"
PA=-81°
0.23"x0.16",
PA=91°
0.23"x0.16",
PA=91°
0.22"x0.15",
PA=90°
0.22"x0.15",
PA=90°
0.22"x0.15",
PA=90°
0.18"x0.16",
PA=80°
0.18"x0.16",
PA=80°
0.17"x0.16",
PA=74°
0.17"x0.16",
PA=75°
0.17"x0.16",
PA=72°
329.063-330.937
330.826-332.700
1.13
341.448-341.917
(CH3CCH J=19-18)
342.263-343.196
0.282
0.98
349.161-349.626
0.25
349.419-349.474
(CH3CN J=19-18)
362.018-362.073
(DCN J=4-3)
362.602-362.657
(HNC J=4-3)
361.948-363.815
0.0625
0.0625
0.0625
1.0
342.363-344.136
1.13
344.063-345.936
1.13
354.477-354.534
(HCN J=4-3)
354.260-356.125
354.677- 354.737
(HC3N J=39-38)
0.061
1.95
0.061
SB
C
July
2:23
22
9.3508 319° 49
B August
9.7504 230° 49
18 23:48
A August
19 1:16
9.7512 230° 49
1 Geocentric distance 2 Polar angle of beam major axis
Entries in bold face indicate data used for wind measurements
16
2) Doppler wind measurements
Wind measurements were obtained from inspection of Doppler shifts on each individual
spectra in the maps. For this, we performed Gaussian fits (the sum of a Gaussian and a 1st
or 2nd degree polynomial) of the line profiles in their central parts, over ranges depending
on the linewidth for each species. The fitting interval was ± [1.875 MHz] for HNC and HC3N
and ± [3.75 MHz] for DCN. For HCN, which shows hyperfine structure, the spectrum was
fit over ± [3.6 MHz] from the main component at 354.5055 GHz, by the sum of three
Gaussians, in order to independently determine Doppler shifts for both the main
component and for the two satellite lines at 354.5039 and 354.5075 GHz. For CH3CN, we
performed separate fits for the three lines at 349.427, 349.447 and 349.454 GHz,
considering each time a ± [2.5 MHz] interval. Results were very similar for each line, and
then averaged over the three lines. For CH3CCH, due to lower line contrasts, we first
created at each beam position a fictitious "total" line, by realigning the spectra
successively on the five strongest lines of the 341.7 GHz multiplet (i.e., the 341.637,
341.682, 341.715, 341.735, and 341.741 GHz lines), and averaging them with weights
proportional to the individual line strengths. The resulting "total" line was then fit for its
position over a ± [8.5 MHz] interval. Wind measurements were obtained in this manner
at all points of the map where the spectra showed sufficient emission signals for the
Gaussian fits to be meaningful. Depending on the species, this was possible when the line
signal exceeded 10 to 35 % of its maximum value on the map. As a result, wind
measurements are usually reliable on a more or less extended region centered on the
limb, an exception being CH3CN for which emission is also strongly detected at all nadir
(disk-intercepting) beam positions, enabling wind maps on the entire Titan disk. For HCN,
we also obtained some wind measurements within the disk using the main component of
the hyperfine triplet, that appears in absorption against the disk (Supplementary Figure
2). To attribute error bars on the winds, we used a simple Monte-Carlo approach48.
Residuals between the observed spectra and the Gaussian fits were used to estimate a rms
noise level. Random noise at this level was then superimposed on the Gaussian model,
which was then refit. Repeating this procedure 300 times, the dispersion in the fitted
positions provided an estimate of the error bars on the LOS wind measurements. In the
case of CH3CN, the error bars were further decreased by averaging the results on the three
components. Wind maps for the 6 species are shown in Fig. 2. Typical error bars for
measurements at/beyond the limb are 12-20 m/s on HNC, 6-8 m/s on HC3N, CH3CN, and
the HCN main line, 10-12 m/s for the two HCN satellite lines, and 30-50 m/s on DCN and
CH3CCH. Errors bars for LOS winds based on the HCN main line within the disk are ~20
m/s.
3) Temperature and abundance retrievals
Temperature profile information was derived from modelling of the CO (3-2) and HCN (4-
3) lines and their isotopic variants, using a home-made inversion code35 and based on
optimal estimation methods principles36-37. For the purpose of this task, which is mainly
to assign altitude levels to the wind measurements, we focused on the equatorial region,
leaving the study of the thermal and compositional field for future work. For the a priori,
we used an equatorial profile derived from Cassini/CIRS measurements in May 2016
(T119 flyby), extended above 700 km at a 153 K temperature, typical of the upper
atmosphere according to Cassini / INMS measurements, and shown in Supplementary
17
Figure 1 (black line). We first used the 12CO (3-2) and 13CO (3-2) lines observed at Equator
and near the central meridian to retrieve the CO mixing ratio (48±1 ppm) and an
intermediate-step thermal profile. The latter (red line in Supplementary Figure 1)
deviates from the a priori only below ~350 km, marking the top of the sensitivity layer to
CO. HCN probes much higher altitudes, however (up to ~1000 km), but the difficulty is
that the temperature profile and the HCN profile are poorly constrained above the top of
the sensitivity range of CIRS (~520 km). Thus in a second step, keeping the temperature
profile fixed below 350 km, we performed a simultaneous inversion of the HCN (4-3) and
H13CN (4-3) lines in terms of temperature and HCN profiles. We assumed a telluric-like
HCN / H13CN ratio of 89, consistent with other measurements of carbon isotopes in Titan's
atmosphere, and took the HCN profile from [39] (extended above 400 km to reach HCN =
10-4 at 800 km and above) as a priori. When doing so, to alleviate the ambiguity between
abundance and temperature effects in the core of the HCN (4-3) line, we inverted
simultaneously the equatorial HCN (4-3) and H13CN (4-3) lines near the central meridian
and at the limb. Indeed, the characteristic absorption / emission appearance of the HCN
hyperfine structure in nadir / limb geometry provides complementary information on
temperature and abundance sensitivity. In practice, we used the west limb spectrum
centered at RA, DEC = (0", -0.42") from Titan center, i.e. 400 km above Titan's solid
surface limb, correcting for the wind Doppler shift before performing the inversion.
The best fit equatorial thermal profile retrieved in this manner is shown in
Supplementary Figure 1 (blue line), and the associated fits of the CO, HCN and H13CN lines
in nadir and limb geometry are shown in Supplementary Figure 2. The temperature
profile shows a marked mesopause at 140 K near 800 km. The occurrence of such a cold
temperature somewhere above the stratopause is in fact directly implied by the depth of
the main component of the HCN (4-3) hyperfine structure, with a brightness temperature
of 143 K in the line core in nadir geometry. Note also that the best fit temperature / HCN
profile permits a good match of the HCN v2=1 emission at 354.460 GHz (Supplementary
Figures 2 and 3). The retrieved profile is also warmer than the a priori (CIRS) profile by
up to ~3 K at 250 km and colder by ~6 K at 400-600 km. Understanding of these
differences is deferred to future work, but we note for the time being that differences
between CIRS- and ALMA CO-derived thermal profiles of up to 5 K below 300 K have been
noted40.
18
1200
1200
1200
1200
1200
1200
1200
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
A priori (Cassini T119, latitude = 1 N)
After CO inversion
After CO inversion
After CO inversion
After CO inversion
After CO inversion
After CO inversion
After CO inversion
1000
1000
1000
1000
1000
1000
1000
After CO and HCN inversion
After CO and HCN inversion
After CO and HCN inversion
After CO and HCN inversion
After CO and HCN inversion
After CO and HCN inversion
After CO and HCN inversion
)
)
)
)
)
)
)
m
m
m
m
m
m
m
k
k
k
k
k
k
k
(
(
(
(
(
(
(
e
e
e
e
e
e
e
d
d
d
d
d
d
d
u
u
u
u
u
u
u
t
t
t
t
t
t
t
i
i
i
i
i
i
i
t
t
t
t
t
t
t
l
l
l
l
l
l
l
A
A
A
A
A
A
A
800
800
800
800
800
800
800
600
600
600
600
600
600
600
400
400
400
400
400
400
400
200
200
200
200
200
200
200
0
0
0
0
0
0
0
60
60
60
60
60
60
60
80
80
80
80
80
80
80
100
100
100
100
100
100
100
120
120
120
120
120
120
120
140
140
140
140
140
140
140
160
160
160
160
160
160
160
180
180
180
180
180
180
180
200
200
200
200
200
200
200
Temperature (K)
Temperature (K)
Temperature (K)
Temperature (K)
Temperature (K)
Temperature (K)
Temperature (K)
Supplementary Figure 1. Thermal profiles. Equatorial thermal profiles used in this
study. Black: a priori profile, based on Cassini/CIRS measurements. Error bars indicate
the adopted a priori errors (see text). They are shown in dotted lines above the region
constrained by Cassini/CIRS. Red: temperature profile retrieved from inversion of 12CO
and 13CO. Blue: thermal profile retrieved from inversion of the HCN and H13CN limb and
nadir data, using the red profile as a priori. Error bars on the retrieved profile include the
propagation of the a priori errors on temperature and HCN (see text). This profile is then
used to determine other molecular profiles and calculate contribution functions.
Error bars on the retrieved HCN/temperature retrievals are a combination of (i)
uncertainties due to measurement noise and (ii) uncertainties due to the a priori used in
the inversion process. Noise-related uncertainties were calculated using the formalism
described in Ref. [37] (their eq. 23 and 24). The associated temperature errors are
formally very small (0.2-0.3 K) in relation to the high S/N and high spectral-resolution of
the data.
19
)
)
m
m
a
a
e
e
b
b
/
/
y
y
J
J
(
(
x
x
u
u
F
F
l
l
)
)
m
m
a
a
e
e
b
b
/
/
y
y
J
J
(
(
x
x
u
u
F
F
l
l
0.5
0.5
0.45
0.45
0.4
0.4
0.35
0.35
0.3
0.3
0.25
0.25
0.2
0.2
0.5
0.5
0.45
0.45
0.4
0.4
0.35
0.35
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
0.5
0.5
0.49
0.49
0.48
0.48
0.47
0.47
0.46
0.46
0.45
0.45
0.44
0.44
0.43
0.43
0.42
0.42
0.41
0.41
0.4
0.4
345.76 345.8
345.76 345.8
Nadir
Nadir
H13CN, HC3N, CO
H13CN, HC3N, CO
345.3
345.3
345.5
345.5
345.7
345.7
345.9
345.9
Frequency (GHz)
Frequency (GHz)
0.52
0.52
0.5
0.5
0.48
0.48
0.46
0.46
0.44
0.44
0.42
0.42
0.4
0.4
0.38
0.38
0.36
0.36
354.5
354.5
354.52
354.52
Nadir
Nadir
HCN, HC3N
HCN, HC3N
354.4 354.6 354.8
354.4 354.6 354.8
355
355
Frequency (GHz)
Frequency (GHz)
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
0.1
0.1
0.05
0.05
0.4
0.4
0.35
0.35
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
0.1
0.1
0.05
0.05
0.26
0.26
0.24
0.24
0.22
0.22
0.2
0.2
0.18
0.18
0.16
0.16
0.14
0.14
345.76 345.8
345.76 345.8
Limb
Limb
H13CN, HC3N, CO
H13CN, HC3N, CO
345.3
345.3
345.5
345.5
345.7
345.7
345.9
345.9
Frequency (GHz)
Frequency (GHz)
Limb
Limb
HCN, HC3N
HCN, HC3N
0.4
0.4
0.35
0.35
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
354.5
354.5
354.52
354.52
354.4
354.4
354.8
354.8
354.6
354.6
Frequency (GHz)
Frequency (GHz)
355
355
Supplementary Figure 2. Spectral fits for CO and HCN. Fits of the CO (3-2), HCN (4-3),
and H13CN (4-3) lines at the equator and in nadir and limb geometries. The insets show
the cores of CO (3-2) and HCN (4-3). Note also, mostly visible in limb geometry, the HCN
v2 line at 354.460 GHz and the HC3N J=38-37 and J=39-38 lines at 345.609 and 354.697
GHz. The associated retrieved profile for HCN is shown in Fig. 3.
20
)
)
m
m
a
a
e
e
b
b
/
/
y
y
J
J
(
(
x
x
u
u
F
F
l
l
)
)
m
m
a
a
e
e
b
b
/
/
y
y
J
J
(
(
x
x
u
u
F
F
l
l
0.32
0.32
0.3
0.3
0.28
0.28
0.26
0.26
0.24
0.24
0.22
0.22
0.2
0.2
0.18
0.18
0.16
0.16
0.14
0.14
0.12
0.12
0.15
0.15
0.13
0.13
0.11
0.11
0.09
0.09
0.07
0.07
0.05
0.05
0.03
0.03
0.01
0.01
Limb
Limb
HCN v2, HCN
HCN v2, HCN
Limb
Limb
CH3CN
CH3CN
0.3
0.3
0.25
0.25
0.2
0.2
0.15
0.15
0.1
0.1
0.05
0.05
354.4
354.4
354.45
354.45
354.5
354.5
354.55
354.55
349.2
349.2
349.3
349.3
349.4
349.4
349.5
349.5
Frequency (GHz)
Frequency (GHz)
Frequency (GHz)
Frequency (GHz)
Limb
Limb
DCN
DCN
Limb
Limb
CH3CCH
CH3CCH
0.15
0.15
0.145
0.145
0.14
0.14
0.135
0.135
0.13
0.13
0.125
0.125
0.12
0.12
0.115
0.115
0.11
0.11
362.03
362.03
362.04
362.04
362.05
362.05
362.06
362.06
341.5
341.5
341.6
341.6
341.7
341.7
341.8
341.8
Frequency (GHz)
Frequency (GHz)
Frequency (GHz)
Frequency (GHz)
Supplementary Figure 3. Spectral fits for other species. Zoom of the HCN v2 line at
354.460 GHz and fits of CH3CN, DCN, and CH3CCH equatorial limb data. The associated
retrieved profiles are shown in Fig. 3.
To evaluate uncertainties due to the a priori, we repeated the temperature/HCN inversion
process by using several a priori temperature and HCN profiles. For this, the first step was
to construct a realistic range of initial temperatures profiles. As indicated above, we
nominally used an equatorial profile derived from Cassini/CIRS measurements45 in May
2016. The formal uncertainty on this profile is < 1 K over 7.5 mbar - 0.025 mbar (100-350
km), increasing to ~ 2 K over 5 x 10-3 -- 1 x 10-3 mbar (430 - 520 km). As there are further
systematic uncertainties on the CIRS-retrieved profile (e.g. related to the precise position
of the CIRS detectors on the sky) at the level of ~1 K over 250-350 km, we conservatively
adopted uncertainties increasing from 1 K at 100 km to 4 K at 630 km. Above this altitude,
we adopted an a priori error that increases as log(pressure) to ~20 K at 1000 km and
above. This choice was guided by (i) the observed dispersion in the INMS isothermal
temperature measurements at 1050 -1500 km, which overall range from 112 K to 175 K
around a mean temperature of 150 K over this altitude range (Ref. [20], Fig. 11), and by
(ii) a similar dispersion in the few available temperature profiles from Cassini / UVIS
(Ref.[42], Fig. 29). We note finally that temperatures at 1000 km from UVIS and HASI
cluster around 180 K, vs 140 K for INMS; therefore, adopting a 153±20 K a priori
temperature at 1000 km is reasonable. These a priori errors defined a set of three initial
temperatures profiles for which the combined CO/HCN retrievals were performed
(Supplementary Figure 1, black line with error bars).
21
We also studied the impact of using different a priori profiles for HCN. In practice, we used
(a) HCN profiles enhanced or depleted from the nominal a priori39 by constant factor-of-
3 (b) an alternative a priori profile in which the HCN kept increasing with altitude up to
5x10-4 at 1000 km (matching the INMS41 and UVIS42 measurements near this altitude),
allowing for a further factor-of-2 uncertainty in this profile. Altogether this provided six a
priori HCN profiles which were also used (along with the nominal a priori temperature)
for the inversion.
Supplementary Figure 4. Sensitivity of retrievals to a priori temperature and HCN
conditions. Retrieved temperature and HCN profiles for a suite of a priori temperature
profiles (shown as the dashed lines in panel a) and a priori HCN profiles (dashed lines in
panel d). The retrieved profiles are shown as solid lines with color codes corresponding
to the a priori. The HCN (resp. temperature) a priori is also shown in panels (b) (resp. (c))
as a grey dashed line.
These modelling variants yielded a suite of solution profiles, whose dispersion is a good
indicator on the actual error bars on the retrieved temperature and HCN profile
(Supplementary Figure 4). Beyond the well-known characteristics of Titan's middle
atmosphere thermal structure such as the stratopause near 270 km, robust new features
of all retrieved profiles are (i) the existence of a quasi-isothermal region over 450-600 km
(already apparent in the CIRS-based a priori) and (ii) the existence of a ~140 K
mesopause, located at 800±50 km altitude depending on the a priori. Error bars on the
retrieved temperatures remain below 2-3 K up to 800 km; above this level they increase
with the progressive lack of sensitivity of the measurements to ultimately reflect the a
priori errors. The HCN profile (Fig. 3) is also very well defined (to within ~10-20 %) over
100-850 km, before error bars progressively increase. The robustness of the retrieved
22
HCN in previous ALMA observations to even factor >10 changes in the a priori was noted
by [44] (their Fig. 2). The nominal HCN mixing profile we determine reaches a 6.3 x 10-4
mixing ratio at 1000 km, in agreement with results from both Cassini/INMS41 and
Cassini/UVIS42, respectively above and below 950 km (see Fig. 3).
The retrieved thermal profile was then used to derive the other molecular profiles, again
for the equatorial region. The DCN profile was taken as the retrieved HCN profile,
multiplied by a constant DCN/HCN ratio of (2.5±0.2) x 10-4 [44], i.e. assuming that the
DCN profile follows that of HCN at all altitudes. CH3CN and CH3CCH were inverted using a
priori profiles respectively taken from ground-based data39 (extended to 10-5 at 1000 km
and above) and CIRS observations45. For these three species, best fits of the (west) limb
equatorial emission are presented in Supplementary Figure 3. Error bars on CH3CN and
CH3CCH reflect the combined effects of noise, temperature profile uncertainty and a priori
errors. They were estimated by allowing for factor-of-2 variations in the a priori, and
using also the warmest / coldest temperature profile when performing the inversion.
The HNC J=4-3 line exhibits a purely Doppler-broadened profile (Supplementary Figure
5), and as such contains limited information on the HNC vertical profile, except for the fact
that this precludes HNC to be present at too deep atmospheric levels. Based on HNC J=6-
5, Ref. [34] found that for constant mixing profiles above some level, the cutoff altitude
cannot be lower than 300 km. We refined this conclusion by testing photochemical
profiles14 for HNC. As the spatial distribution of HNC J=4-3 shows only mild latitudinal
variations, we averaged the data over a concentric ring (0.285"-0.475", i.e. 2000-3350 km
from Titan center) to improve S/N. We found (Supplementary Figure 5, top right) that
several of the HNC models from Ref. [14], in particular the nominal one, overestimate the
amount of HNC in the lower stratosphere, producing unobserved spectral wings. In
contrast, the model in which the production of HNC by Galactic Cosmic Ray impact is
omitted is essentially in agreement with the observed spectrum, requiring only a modest
rescaling by a factor 1.15. Nonetheless, this line shape-based approach does not constrain
the distribution of HNC at altitudes above the pressure-broadening levels. The latter can
be determined by inspecting the observed variation of the line-integrated intensities with
radial distance. Specifically, we compared the latter with model calculations for constant
HNC profiles above some cutoff altitudes, using the nominal retrieved temperature profile
(blue profile in Supplementary Figure 1). Supplementary Figure 5 (top left) shows the
result, with cutoff altitudes of 600, 800, 1000, 1200 km, considering together all the HNC
data. Too low (resp. too high) cutoff altitudes lead the models to predict an emission
maximum at too small (resp. too large) radial distances. Part of the dispersion of the
observational points in this panel is due to the spatial variation of the HNC emission.
Repeating the exercise, both by (i) generating synthetic datasets with similar dispersion
as the observed and (ii) splitting the HNC measurements in angular bins of 30° across the
disk, provides a robustness test of the method and indicates that the bulk of HNC is located
above 870±40 km altitude, implying that HNC is mostly a thermospheric species.
Furthermore, Supplementary Figure 5 (top left) shows that the photochemical model
from Ref. [14] without GCR also provides the correct radial dependence for the HNC
emission. As this profile has more physical basis than a piecewise distribution, we adopt
it as nominal profile, allowing for a vertical shift of ±40 km in its altitude scale (see Fig. 3).
Using extreme temperature profiles does not change the altitude distribution of HNC by
more than 10 km.
23
The situation is similar for the HC3N J=39-38 emission at 354.697 GHz, which -- except in
the high-northern and southern latitude regions where it is strongly enhanced -- does not
show Lorentzian wings. We performed the same study as for HNC on the radial distance
variation of the HC3N emission (using the slightly stronger J=38-37 line at 345.609 GHz
observed along with CO (3-2)), restricting ourselves to points of the maps having a
declination offset within [± 0.2"], i.e. [± 0.55 Titan's radius] from disk center, in order to
focus on the low-latitude regions. Fitting models in which HC3N is assumed to be
uniformly mixed above some cutoff altitudes and again refitting synthetic datasets with
an appropriate noise level indicates a best fit cutoff altitude of 630±30 km
(Supplementary Figure 5, bottom left), implying that HC3N too is dominantly present in
Titan's upper atmosphere, albeit less so than HNC. This is not surprising given that HC3N,
as measured by CIRS12-13 has also a measurable abundance in Titan's stratosphere and
lower mesosphere, with mixing ratios increasing from 6x10-11 -- 1x10-7 over 150-450 km.
Initializing the HC3N profile at these values and a large abundance (2x10-4) at z > 800 km,
we determined from inversion an equatorial HC3N profile fitting both the shape of the
J=39-38 line at equatorial west limb and the radial variation of the line-integrated J=38-
37 lines fluxes. In contrast (Supplementary Figure 5, bottom right), the nominal HC3N
profile from Ref. [14] does not fit any either of the two observables, indicating that this
profile is overabundant at lower altitudes and under-abundant in the upper atmosphere.
Our HC3N solution profile has a 1000 km abundance as high as ~5 x 10-5. This agrees
rather well with the global value reported from Cassini/INMS46 (3.2 x 10-5 at 1025 km,
after correction for possible effects of wall adsorption/desorption), and with the
Cassini/UVIS HC3N abundance over 850-950 km [42]. We reiterate that our conclusions
pertain to the equatorial HC3N profile only.
24
Supplementary Figure 5. Determination of the vertical profile of HNC and HC3N
from combination of line shape (right panels) and radial dependence (left panels)
information. Left panels: line-integrated fluxes (Jy-MHz per beam) for HNC (4-3) and
HC3N (38-37) as a function of distance (in Titan radius) from Titan's center. For HNC, all
points on the map are considered, while for HC3N, only low-latitude regions are
considered (see text). These radial profiles are compared both with piecewise profiles
(HNC or HC3N assumed to be uniformly mixed above some cutoff altitudes) and with
inverted (HC3N) or physically-based profiles (models from Dobrijevic et al.14 for HNC and
HC3N). HNC: For HNC, the piece-wise profiles have the following abundances: 1.6x10-7
above 600 km, 5.6x10-6 above 800 km, 1.25x10-4 above 1000 km and 2.0x10-3 above 1200
km. The models are shifted upwards by 0.05 units for clarity, and the "noise" in the models
results from the non-circular beam. The best fit for piecewise models is achieved for q HNC
= 1.7x10-5 above 870 km. The HNC profile in which the dissociation by Galactic Cosmic
Ray (GCR) is turned off (black profile in Fig. 26 of Ref. [14]), when multiplied by a modest
factor of 1.15, also provides an excellent match to the radial dependence of the emission.
Top right panel: HNC J = (39-38) line, extracted on an annulus ranging from 2000-3350
km from Titan center. The previous model with no GCR contribution matches the line
shape; in contrast, other models from Ref. [14] (red and blue profiles in their Fig. 26) have
too much HNC in the lower mesosphere below 600 km and produce unobserved line-
wings. HC3N: For HC3N, the piecewise profiles have the following abundances: 1.05x10-7
above 400 km, 4.4x10-6 above 600 km, and 1.4x10-4 above 800 km. The models are shifted
upwards by 0.07 units for clarity. The best fit for piecewise models is achieved for q HC3N
= 7x10-6 above 630 km. Due to the selection criterion for the HC3N data, the associated
profile pertains to the equatorial region only. Bottom right panel. HC3N J=39-38 line on
the equatorial west limb, compared to best fit model (light blue) resulting from inversion
of the line profile and including constraints from the radial dependence of the line-
integrated HC3N (38-37) fluxes at low latitude. In contrast, the nominal HC3N profile of
25
Ref. [14] (their Fig. 10, red line) fails at matching both observables (yellow symbols in the
two bottom panels), indicating that that profile has too much HC3N at low altitudes and
not enough at high altitudes.
Figure 3 summarizes the retrieved profiles for all six molecules used to probe the winds.
Error bars are indicated in the range where information is available. For HNC and HC3N,
uncertainties are best expressed as a vertical shift of ±40 km and ±30 km in the profiles,
for reasons explained above. These profiles are compared to characteristic profiles locally
derived from Cassini/UVIS at 6°S (T41I, Ref. [42]), INMS41,46, and CIRS. For the latter, we
show HCN and CH3CCH profiles at 9°S from May 2016 (T119, Ref. [45]); for HC3N, not
clearly detected in the latter observations, we use instead a 5°N profile from Jan. 2007
(T23, Ref. [47]). Also shown in Fig. 3 are disk-averaged HCN and HC3N profiles derived
from 2012-2015 ALMA observations ([33], their Fig. 10). Figure 3 highlights the overall
agreement between these profiles, giving in particular great confidence in our HCN and
HC3N profiles in the middle/upper atmosphere. A noteworthy difference is that "local"
molecular profiles (from Cassini CIRS and UVIS) show small-scale vertical variations (see
e.g. Figs. 3-4 of Ref. [47], and Fig. 26 of Ref. [42]), not predicted in photochemical models14
and presumably of dynamical origin; those are particularly visible in Fig. 3 for the HCN
and HC3N profiles derived from UVIS.
4) Wind structure and wind models
In Supplementary Figure 6, the line-of-sight wind velocities are plotted as a function of
the "algebraic" projected distance to Titan's axis -- by definition, the latter is positive (resp.
negative) sky-east (resp. west) of the polar axis, and the distance is expressed in units of
Titan's "effective" radius, i.e. the (species-dependent) radial distance at the altitude of the
wind contribution function (WCF)-weighted mean probed altitude. (Given the strong
variation of the HCN WCF between nadir and limb, HCN LOS winds within the disk are not
included). In this type of representation (and ignoring for now beam convolution effects),
a solid-body rotation appears as a straight line with slope equal to veq cos(β), where β =
26°N is the sub-Earth latitude. (In contrast, such plots do not permit one to investigate
latitudinal dependences). Best fit and error bars for veq were obtained from error-bar
weighted fits of the individual measurements. In Supplementary Figure 6, LOS winds from
different molecules are compared two-by-two, and linear fits are over-plotted with best
fit veq values indicated. This figure highlights several facts: (i) wind values from CH3CN,
CH3CCH, and DCN show excellent consistency (despite the larger uncertainties on the
latter two species), both in their structure and in the veq value (191-206 m/s) (ii) HNC
winds are clearly different in speed and structure; although a solid-body rotation model
leads to a best fit veq = 297 m/s, they do not align well on a straight line, instead showing
a deviation from the solid body rotation to higher velocities at large distances, i.e. an
equatorial jet (iii) The veq values for the HCN main line and HC3N winds are virtually
identical (287-290 m/s), and the structures seen in their maps seem intermediate
between those of CH3CN and HNC (although HCN exhibits a stronger "tongue" at large
distances than HC3N) (iv) although the HCN main line nominally probes the atmosphere
26
~200 km higher than the satellite lines (see Fig. 3), the associated veq values are only
marginally different. The last panel of Supplementary Figure 6 shows the variation of the
fitted equatorial velocity with altitude, as indicated by the WCF for each molecule. These
findings are consistent with the three altitude regions outlined above and with a general
increase of winds with altitude, with a marked gradient over 350-600 km. Still, the vertical
shear du/dz (using u = veq) is small enough that the Richardson number Ri = N2 / (du/dz)2,
where N is the Brunt -- Väisälä frequency, rises from ~20 at 300 km to ~1000 at 1000 km.
This is safely above the threshold Ri < ¼ for turbulence.
The above description is approximate, especially because it ignores beam-convolution
effects. In addition, the data indicate more complex structure than solid-body rotation,
with a progressive onset of the jet regime above ~600 km. To estimate the wind regime
indicated by the data, we simulated the expected LOS Doppler shifts, testing various cases
for the latitudinal distribution of winds. The expected Doppler shifts were calculated
locally by considering a globe with radius defined by the mean probed altitude and
projecting the zonal winds accordingly, accounting for the viewing geometry with sub-
earth latitude at 26°N. In doing so, we also included the minor effect of Titan's solid body
rotation with orbital period of 15.95 days (giving e.g. a 16 m/s equatorial speed at 990
km altitude, the nominal altitude probed by HNC). Only purely zonal winds were
considered, e.g. no meridional components were included. The modelled local LOS
Doppler shifts were then weighted by the local line intensity -- i.e. the line emission signal
integrated over the spectral interval used to measure the winds in the data -- and finally
convolved by the instrumental beam. We performed this exercise for winds measured in
CH3CN, HNC and HC3N, noting that this is still an approximate way of modelling the data.
Indeed, given the vertical extent of the wind contribution functions, especially for HC3N
and HCN, the only completely rigorous approach would consist of testing 2D (latitude,
altitude) wind fields and generate the associated synthetic line profiles, incorporating the
altitude-dependent Doppler shifts in the absorption coefficients before performing the
radiative transfer. Such synthetic profiles would finally be refit for Doppler shifts in the
same manner as the actual data. This complex task, particularly needed for HC3N and HCN,
is left for the future. With this caveat in mind, Supplementary Figure 7 shows synthetic
wind models for CH3CN, HNC and HC3N that provide a satisfactory match to the data
shown in Fig. 2. As is clear from the first panel of Supplementary Figure 7, the zonal wind
magnitude and its latitudinal dependence are different between the three species,
indicative of an evolution of the wind regime with altitude.
Meridional winds are not obviously present in the data. Assuming equator-to-pole or
pole-to-pole circulation, a meridional circulation would be best visible at mid-latitudes
along the central meridian. Only the CH3CN data are suited for that, and based on the wind
contours in Fig. 2, we estimate an upper limit of ~25 m/s for such winds. Similarly, day-
to-night flows would be best captured from Doppler shifts in the polar region (e.g.
receding winds at the poles for SSAS winds). Our data do not give any consistent evidence
for this either, with an upper limit estimated to ~30 m/s.
27
Supplementary Figure 6. Line-of-sight wind velocities as a function of projected
distance from polar axis. The distance is expressed in units of "Titan effective radius",
defined as the radial distance at the WCF-weighted mean probed altitude (e.g. 3565 km
for HNC). By convention, the distance is positive (resp. negative) for points located sky-
east (resp. west) from the polar axis. Straight lines show simple linear fits to the data
assuming solid-body rotation (see text) with the corresponding equatorial velocities
indicated. The fit for CH3CN (black line) is repeated on all panels for comparison. The last
panel shows the variation of the fitted equatorial velocity with altitude. In that panel, the
color code is used consistently with the other panels, and the vertical error bars indicate
the extent of the wind contribution functions (see Fig. 3).
28
Supplementary Figure 7. Modelled wind structure. Simulated Doppler shifts for HNC,
HC3N and CH3CN, using the respective latitude profiles of the zonal wind shown in the first
panel. Local line-of-sight winds are weighted by line intensities and convolved by the
instrumental beam. Titan's solid-body rotation is included. These figures approximate the
observed maps for these species (see Fig. 2).
29
|
0908.1672 | 2 | 0908 | 2009-08-21T19:34:38 | HAT-P-7: A Retrograde or Polar Orbit, and a Third Body | [
"astro-ph.EP",
"astro-ph.SR"
] | We show that the exoplanet HAT-P-7b has an extremely tilted orbit, with a true angle of at least 86 degrees with respect to its parent star's equatorial plane, and a strong possibility of retrograde motion. We also report evidence for an additional planet or companion star. The evidence for the unparalleled orbit and the third body is based on precise observations of the star's apparent radial velocity. The anomalous radial velocity due to rotation (the Rossiter-McLaughlin effect) was found to be a blueshift during the first half of the transit and a redshift during the second half, an inversion of the usual pattern, implying that the angle between the sky-projected orbital and stellar angular momentum vectors is 182.5 +/- 9.4 degreees. The third body is implicated by excess radial-velocity variation of the host star over 2 yr. Some possible explanations for the tilted orbit are a close encounter with another planet, the Kozai effect, and resonant capture by an inward-migrating outer planet. | astro-ph.EP | astro-ph |
DRAFT VERSION NOVEMBER 2, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
HAT-P-7: A RETROGRADE OR POLAR ORBIT, AND A THIRD BODY
ANDREW W. HOWARD4,5, GEOFFREY W. MARCY4, IAN J. CROSSFIELD6, MATTHEW J. HOLMAN7
JOSHUA N. WINN1, JOHN ASHER JOHNSON2,3, SIMON ALBRECHT1,
Draft version November 2, 2018
ABSTRACT
We show that the exoplanet HAT-P-7b has an extremely tilted orbit, with a true angle of at least 86◦ with
respect to its parent star's equatorial plane, and a strong possibility of retrograde motion. We also report
evidence for an additional planet or companion star. The evidence for the unparalleled orbit and the third body
is based on precise observations of the star's apparent radial velocity. The anomalous radial velocity due to
rotation (the Rossiter-McLaughlin effect) was found to be a blueshift during the first half of the transit and
a redshift during the second half, an inversion of the usual pattern, implying that the angle between the sky-
projected orbital and stellar angular momentum vectors is 182.◦5 ± 9.◦4. The third body is implicated by excess
radial-velocity variation of the host star over 2 yr. Some possible explanations for the tilted orbit are a close
encounter with another planet, the Kozai effect, and resonant capture by an inward-migrating outer planet.
Subject headings: planetary systems -- planetary systems:
formation -- stars: individual (HAT-P-7) --
stars: rotation
1. INTRODUCTION
In the Solar system, the planetary orbits are well-aligned
and prograde, revolving in the same direction as the rotation
of the Sun. This fact inspired the "nebular hypothesis" that the
Sun and planets formed from a single spinning disk (Laplace
1796). One might also expect exoplanetary orbits to be well-
aligned with their parent stars, and indeed this is true of most
systems for which it has been possible to compare the direc-
tions of orbital motion and stellar rotation (Fabrycky & Winn
2009, Le Bouquin et al. 2009). However, there are at least 3
exoplanets for which the orbit is tilted by a larger angle than
any of the planets in the Solar system: XO-3b (Hébrard et
al. 2008, Winn et al. 2009a), HD 80606b (Moutou et al. 2009,
Pont et al. 2009, Winn et al. 2009b), and WASP-14b (Johnson
et al. 2009).
Still, all of those systems are consistent with prograde or-
bits, with the largest minimum angle between the stellar-
rotational and orbital angular momentum vectors of about
37◦, for XO-3b (Winn et al. 2009a). The reason why only the
minimum angle is known is that the evidence for misalign-
ment is based on the eponymous effect of Rossiter (1924) and
McLaughlin (1924), an anomalous Doppler shift observed
during planetary transits that is sensitive only to the angle be-
tween the sky projections of the two vectors. The true spin-
orbit angle may be larger, depending on the unknown inclina-
tion angle of the stellar rotation axis with respect to the line
of sight.
In this Letter we present evidence of a very large spin-orbit
misalignment for HAT-P-7b, a planet of mass 1.8 MJup and
1 Department of Physics, and Kavli Institute for Astrophysics and Space
Research, Massachusetts Institute of Technology, Cambridge, MA 02139
2 Institute for Astronomy, University of Hawaii, Honolulu, HI 96822
3 NSF Astronomy and Astrophysics Postdoctoral Fellow
4 Department of Astronomy, University of California, Mail Code 3411,
Berkeley, CA 94720
geles, CA 90095
bridge, MA 02138
5 Townes Postdoctoral Fellow, Space Sciences Laboratory, University of
California, Berkeley, CA 94720
6 Department of Physics and Astronomy, University of California, Los An-
7 Harvard-Smithsonian Center for Astrophysics, 60 Garden St., Cam-
radius 1.4 RJup in a 2.2-day orbit around an F6V star with
mass 1.5 M⊙ and radius 1.8 R⊙ (Pál et al. 2008). We find the
angle between the sky-projected angular momentum vectors
to be 182.◦5 ± 9.◦4. Furthermore we show that the true angle
ψ between those vectors is likely greater than 86◦, indicating
that the orbit is either retrograde (ψ > 90◦) or nearly polar
(ψ ≈ 90◦). We also present evidence for a third body in the
system, which may be an additional planet or a companion
star. We present spectroscopic data in § 2, photometric data
in § 3, a joint analysis of both types of data in § 4, and a
discussion of the results in § 5.
2. RADIAL VELOCITIES
We observed HAT-P-7 with the High Resolution Spectro-
graph (HIRES) on the Keck I 10m telescope, and the High
Dispersion Spectrograph (HDS) on the Subaru 8m telescope.
The planet's discoverers (Pál et al. 2008; hereafter, P08) ob-
tained 8 HIRES spectra in 2007, to which we add 9 spectra
from 2009. All but one of the HIRES spectra were acquired
outside of transits. Of the 49 HDS spectra, 9 were obtained
on 2009 June 17 and 40 were obtained on 2009 July 1. The
second of these nights spanned a transit.
The instrument settings and observing procedures in both
2007 and 2009 were identical to those used by the California
Planet Search (CPS; Howard et al. 2009). We placed an io-
dine gas absorption cell into the optical path, to calibrate the
instrumental response and wavelength scale. The radial ve-
locity (RV) of each spectrum was measured with respect to an
iodine-free template spectrum, using the algorithm of Butler
et al. (2006) with subsequent improvements. Measurement
errors were estimated from the scatter in the fits to individual
spectral segments spanning a few Angstroms. The RVs are
given in Table 1.
2.1. Evidence for a third body
Fig. 1 shows the RVs over the 2 yr span of the observations.
Fig. 2 shows the RVs as a function of orbital phase, fitted
with 2 different models. The first model is a single Keplerian
orbit, representing the signal of the known planet. The sec-
ond model has an additional parameter γ representing an ex-
tra radial acceleration. The second model gives a better fit to
2
Winn, Johnson, Albrecht, et al. 2009
FIG. 1. -- Long-term radial velocity variation of HAT-P-7. (a) Measured RVs. (b) Residuals (observed -
calculated) be-
tween the data and the best-fitting single-planet model. Light blue and dark blue points are HIRES data from 2007 and 2009,
respectively.
the data, with a root-mean-squared (rms) residual of 7 m s- 1
as compared to 21 m s- 1 for the first model. The RVs from
2009 are systematically redshifted by approximately 40 m s - 1
compared to RVs from 2007, as evident from the residuals
shown in Figs. 1(b) and 2(b). This shift is highly significant,
as the CPS has demonstrated a long-term stability of 2 m s- 1
or better using HIRES and the same reduction codes used here
(Howard et al. 2009).
A transiting planet in a well-aligned prograde orbit would
first pass in front of the blueshifted (approaching) half of the
star, causing an anomalous redshift of the observed starlight.
Then, the planet would cross to the redshifted (receding)
half of the star, causing an anomalous blueshift. In contrast,
Fig. 3(a) shows a blueshift followed by a redshift: an inver-
sion of the effect just described. We may conclude, even with-
out any modeling, that the orbital "north pole" and the stellar
"north pole" point in nearly opposite directions on the sky.
3. PHOTOMETRY
For a quantitative analysis of the RM effect we wanted to
model both the photometric and spectroscopic transit signals.
For this purpose we supplemented the RV data with the most
precise transit light curve available to us, shown in Fig. 3(c).
This light curve is based on observations on UT 2008 Sep 22
in the Sloan i bandpass, with the Fred L. Whipple 1.2m tele-
scope and Keplercam detector, under the auspices of the Tran-
sit Light Curve project (Holman et al. 2006, Winn et al. 2007).
Reduction of the CCD images involved standard procedures
for bias subtraction and flat-field division. Differential aper-
ture photometry was performed for HAT-P-7 and 7 compari-
son stars. No evidence was found for time-correlated noise us-
ing the "time-averaging" method of Pont et al. (2006), as im-
plemented by Winn et al. (2009c). The data shown in Fig. 3(c)
were corrected for differential extinction as explained in § 4.
4. JOINT ANALYSIS
We fitted a model to the photometric and RV data in or-
der to derive quantitative constraints on the angle λ between
the sky projections of the orbital and stellar-rotational angu-
lar momentum vectors. This angle is defined such that λ = 0◦
when the sky-projected vectors are parallel and λ = 180◦ when
they are antiparallel. Our model for the RM effect was based
on the technique of Winn et al. (2005): we simulated spectra
exhibiting the RM effect at various transit phases, and then
measured the apparent RV of the simulated spectra using the
This RV trend is evidence for an additional companion.
Given the limited time coverage of our observations (two clus-
ters of points separated by 2 yr), the data are compatible with
nearly any period longer than a few months. A constant ac-
celeration is the simplest model that fits the excess RV vari-
ability, and under that assumption we may give an order-of-
magnitude relation relating γ to some properties of the com-
panion,
Mc sin ic
a2
c
∼
γ
G
= (0.121 ± 0.014) MJup AU- 2,
(1)
where Mc is the companion mass, ic its orbital inclination rel-
ative to the line of sight, ac its orbital distance, and the numer-
ical value is based on our model-fitting results (see § 4).
2.2. Evidence for a spin-orbit misalignment
Fig. 3(a) shows the RV data spanning the transit, after sub-
tracting the orbital RV as computed with the best-fitting model
including γ. We interpret the "anomalous" RV variation dur-
ing the transit as the Rossiter-McLaughlin (RM) effect, the
asymmetry in the spectral lines due to the partial eclipse of the
rotating photosphere. In the context of eclipsing binary stars,
the RM effect was predicted by Holt (1893) and observed
definitively by Rossiter (1924) and McLaughlin (1924). For
exoplanets, the RM effect was first observed by Queloz et
al. (2000), and its use in assessing spin-orbit alignment has
been expounded by Ohta et al. (2005) and Gaudi & Winn
(2007).
Outlandish orbit of HAT-P-7b
3
FIG. 2. -- Phased radial velocity variation of HAT-P-7. (a) Assuming a single Keplerian orbit. (b) Residuals. (c) With an extra
parameter γ representing a constant radial acceleration. (d) Residuals. The circles are HIRES data (light blue from 2007, dark
blue from 2009), the green triangles are HDS data from 2009 June 17, and the red squares are HDS data from 2009 July 01.
same Doppler code that is used on actual data. This allowed
us to relate the anomalous RV to the parameters and positions
of the star and planet.
The RV model was the sum of the Keplerian RV and the
anomalous RV due to the RM effect. The photometric model
was based on the analytic equation for the flux of a quadrati-
cally limb-darkened disk with a circular obstruction (Mandel
& Agol 2002). As a compromise between fixing the limb-
darkening coefficients u1 and u2 at theoretically calculated
values, and giving them complete freedom, we fixed u1 - u2
at the tabulated value of 0.3846 (Claret 2004) and allowed
u1 + u2 to be a free parameter. We also included a free pa-
rameter for the coefficient of differential airmass extinction
between HAT-P-7 and the ensemble of comparison stars.
We determined the best values of the model parameters and
their 68.3% confidence limits using a Markov Chain Monte
Carlo algorithm, as described in our previous works (see, e.g.,
Winn et al. 2009a). The likelihood function was given by
exp(- χ2/2) with
(cid:20) fi(obs) -
(cid:20) vi(obs) - vi(calc)
fi(calc)
(cid:21)2
(cid:21)2
χ2 =
,
+
Nv
Xi=1
N f
Xi=1
σ f ,i
σv,i
(2)
in a self-explanatory notation, with σ f ,i chosen to be 0.00136,
and σv,i chosen to be the quadrature sum of the RV measure-
ment error and a "stellar jitter" term of 9.3 m s- 1. These
choices led to χ2 = Ndof for the minimum-χ2 model. A Gaus-
sian prior constraint was imposed upon the orbital period
based on the precise measurement of P08.
Table 2 gives the results for the model parameters. In par-
ticular, the result for λ is 182.5 ±9.4 deg, close to antiparallel,
4
Winn, Johnson, Albrecht, et al. 2009
FIG. 3. -- The spectroscopic and photometric transit of HAT-P-7b. (a) The anomalous RV, defined as the output of the Doppler
code minus the orbital RV. We observed a blueshift in the first half of the transit, and a redshift in the second half of the transit,
demonstrating that the sky projections of the orbital and stellar angular momentum vectors point in opposite directions. (b)
Residuals. Red squares are HDS data from 2009 July 1, and blue circles are HIRES data obtained on various nights in 2007 and
2009. (c) The relative flux, observed in the Sloan i band with the FLWO 1.2m telescope and Keplercam. (d) Residuals. In panels
(a) and (b), the gray line shows the best fitting model.
as anticipated from the qualitative discussion of § 2.
5. DISCUSSION
Our finding for λ is strongly suggestive of retrograde mo-
tion, in which the orbital motion and stellar rotation are in
opposite directions. However, it must be remembered that λ
refers to the angle between the sky-projected angular momen-
tum vectors. The true angle ψ between the vectors is given by
(3)
where i and i⋆ are the line-of-sight inclinations of the or-
bital and stellar angular momentum vectors, respectively. Al-
though i is known precisely from the transit data, i⋆ is un-
cos ψ = cosi⋆ cosi + sini⋆ sin icos λ,
known.
Supposing i⋆ to be drawn from an "isotropic" distribu-
tion (uniform in cosi⋆), the data demand that ψ > 86.◦3 with
99.73% confidence. Thus, under this assumption, a retrograde
orbit is strongly favored, although a nearly-polar and barely-
prograde orbit cannot be ruled out.
In fact there is circumstantial evidence that i⋆ is small and
consequently the orbit of HAT-P-7b is nearly polar (ψ ≈
90◦). The star's projected rotation rate is unusually low for
such a hot star: vsin i⋆ = 4.9+1.2
- 0.9 km s- 1 in our model, or
3.8 ± 0.5 km s- 1 based on the line profile analysis of P08;
and Teff = 6350 ± 80 K according to P08.
In the SPOCS
catalog of dwarf stars with well-determined spectroscopic
Outlandish orbit of HAT-P-7b
5
properties (Valenti & Fischer 2005), only 2 of 37 stars with
Teff = 6350 ± 100 K have vsin i⋆ < 4.9 km s- 1.
Based on this catalog, the mean rotation rate v for such hot
stars is about 15 km s- 1. As an alternate approach to con-
straining ψ, we assumed the rotation velocity v is drawn from
a Gaussian distribution with mean 15 km s- 1 and standard de-
viation 3 km s- 1. The result is ψ = 94.6+5.5
- 3.0 deg with 68.3%
confidence, and ψ > 86.◦1 with 99.73% confidence. This anal-
ysis favors nearly-polar and retrograde orbits. However, one
wonders whether HAT-P-7 should be expected to have a "typi-
cal" rotation rate, given the existence of its short-period planet
on a bizarre orbit. Another caveat is that we found the scaled
semimajor axis a/R⋆ to be about 1σ smaller than the find-
ing of P08, suggesting the star is somewhat larger and more
evolved, which would correspond to a slower expected rota-
tion rate.
Determining i⋆ directly may be possible by measuring
and interpreting asteroseismological oscillations (Gizon &
Solanki 2003), or photometric modulations produced by
starspots (see, e.g., Henry & Winn 2008). By good fortune,
HAT-P-7 is in the field of view of the Kepler satellite, which
is capable of precise long-term photometry and may be able
to accomplish these tasks (Borucki et al. 2009).
The extraordinary orbit of HAT-P-7b presents an extreme
case for theories of planet formation and subsequent orbital
evolution. HAT-P-7b is a "hot Jupiter" and presumably mi-
grated inward toward the star after its formation. A prevail-
ing migration theory involves tidal interactions with the proto-
planetary disk, but such interactions would probably not per-
turb the initial coplanarity of the system, and might even bring
the system into closer alignment (Lubow & Ogilvie 2001,
Cresswell et al. 2007). More promising to explain HAT-P-7b
are scenarios involving few-body dynamics, as those scenar-
ios are expected to produce misalignments. In one scenario,
close encounters between planets throw a planet inward,
where its orbit is ultimately shrunk and circularized by tidal
dissipation (Chatterjee et al. 2008, Juri´c & Tremaine 2008).
Another idea is based on the Kozai (1962) effect, whereby the
gravitational force from a distant body on a highly inclined
orbit strongly modulates an inner planet's orbital eccentricity
and inclination (Fabrycky & Tremaine 2007). Recent calcu-
lations showed that a combination of planet-planet scattering,
the Kozai effect, and tidal friction can lead to nearly-circular
retrograde orbits (Nagasawa et al. 2008). A third proposed
scenario involves an inward-migrating outer planet that cap-
tures an inner planet into a mean motion resonance; if the
inner planet avoids being ejected or consumed by the star, it
may be released on a nearly-circular retrograde orbit (Yu &
Tremaine 2001).
The prospect of explaining HAT-P-7b's orbit through few-
body dynamics lends extra importance to measuring the mass
and orbital parameters of the third body.
If it turns out to
be a planet, then HAT-P-7b will be only the second known
case of a transiting planet accompanied by another planet,
the first being HAT-P-13b (Bakos et al. 2009). Such systems
are highly desirable because the unusually precise measure-
ments enabled by transit observations can be used to deter-
mine whether the orbits are coplanar and give clues about the
system's dynamical history (Fabrycky 2009).
Note added after submission. -- Narita et al. (2009) report
independent evidence for a retrograde or polar orbit of HAT-
P-7b, based on Subaru/HDS spectra spanning the transit of
2008 May 30.
We are grateful to Yasushi Suto and Ed Turner for stimu-
lating our interest in this subject; Norio Narita and his team
for sharing their data in advance of publication; Dan Fab-
rycky, Guillaume Hébrard, András Pál, Darin Ragozzine,
Scott Tremaine, Bill Welsh, and the anonymous referee for
helpful comments on the manuscript; Akito Tajitsu, Tae-Soo
Pyo, Mark Everett, Howard Isaacson, and Zach Gazak for as-
sistance with observing; Gáspár Bakos and Joel Hartman for
help obtaining telescope time; Eric Gaidos and Debra Fischer
for trading telescope time on short notice; and Hector Balbon-
tin for hospitality at Las Campanas Observatory where this
manuscript was written.
Some of the data presented herein were obtained at the
W.M. Keck Observatory, which is operated as a scientific
partnership among the California Institute of Technology, the
University of California, and the National Aeronautics and
Space Administration, and was made possible by the gener-
ous financial support of the W.M. Keck Foundation. We ex-
tend special thanks to those of Hawaiian ancestry on whose
sacred mountain of Mauna Kea we are privileged to be guests.
Without their generous hospitality, the Keck observations pre-
sented herein would not have been possible.
J.A.J. grate-
fully acknowledges support from the NSF Astronomy and As-
trophysics Postdoctoral Fellowship program (grant no. AST-
0702821). S.A. acknowledges the support of the Nether-
lands Organisation for Scientific Research (NWO). J.N.W.
gratefully acknowledges support from the NASA Origins pro-
gram through awards NNX09AD36G and NNX09AB33G,
and from an MIT Class of 1942 Career Development Profes-
sorship.
Facilities: Subaru (HDS), Keck:I (HIRES), FLWO:1.2m
(Keplercam)
REFERENCES
Bakos, G. A., et al. 2009, arXiv:0907.3525
Borucki, W. J., et al. 2009, Science, 325, 709
Butler, R. P., et al. 2006, ApJ, 646, 505
Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008, ApJ, 686,
Holt, J. R. 1893, Astronomy and Astro-Physics, 12, 646
Howard, A. W., et al. 2009, ApJ, 696, 75
Johnson, J. A., Winn, J. N., Albrecht, S., Howard, A. W., Marcy, G. W., &
Gazak, J. Z. 2009, arXiv:0907.5204
Juri´c, M., & Tremaine, S. 2008, ApJ, 686, 603
Kozai, Y. 1962, AJ, 67, 591
Laplace, P. S. 1796, Exposition du système du monde (Paris: Cercle-Social)
Le Bouquin, J.-B., Absil, O., Benisty, M., Massi, F., Mérand, A., & Stefl, S.
2009, A&A, 498, L41
Lubow, S. H., & Ogilvie, G. I. 2001, ApJ, 560, 997
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
McLaughlin, D. B. 1924, ApJ, 60, 22
Moutou, C., et al. 2009, A&A, 498, L5
Nagasawa, M., Ida, S., & Bessho, T. 2008, ApJ, 678, 498
Narita, N., Sato, B., Hirano, T., & Tamura, M. 2009, arXiv:0908.1673
Ohta, Y., Taruya, A., & Suto, Y. 2005, ApJ, 622, 1118
580
Claret 2004, A&A, 428, 1001
Cresswell, P., Dirksen, G., Kley, W., & Nelson, R. P. 2007, A&A, 473, 329
Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298
Fabrycky, D. C., & Winn, J. N. 2009, ApJ, 696, 1230
Fabrycky, D. C. 2009, in Proc. IAU Symposium 253, eds. F. Pont, D.D.
Sasselov, & M.J. Holman (Cambridge: CUP), 173
Gaudi, B. S. & Winn, J. N. 2007, ApJ, 655, 550
Gizon, L., & Solanki, S. K. 2003, ApJ, 589, 1009
Hébrard, G., et al. 2008, A&A, 488, 763
Henry, G. W., & Winn, J. N. 2008, AJ, 135, 68
Holman, M. J., et al. 2006, ApJ, 652, 1715
6
Winn, Johnson, Albrecht, et al. 2009
Pál, A., et al. 2008, ApJ, 680, 1450
Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS, 373, 231
Pont, F., et al. 2009, A&A, 502, 695
Queloz, D., Eggenberger, A., Mayor, M., Perrier, C., Beuzit, J. L., Naef, D.,
Sivan, J. P., & Udry, S. 2000, A&A, 359, L13
Rossiter, R. A. 1924, ApJ, 60, 15
Valenti, J. A., & Fischer, D. A. 2005, ApJS, 159, 141
Winn, J. N., et al. 2005, ApJ, 631, 1215
Winn, J. N., Holman, M. J., & Fuentes, C. I. 2007, AJ, 133, 11
Winn, J. N., et al. 2009a, ApJ, 700, 302
Winn, J. N., et al. 2009b, ApJ, in press [arXiv:0907.5205]
Winn, J. N., et al. 2009c, ApJ, 693, 794
Yu, Q., & Tremaine, S. 2001, AJ, 121, 1736
RELATIVE RADIAL VELOCITY MEASUREMENTS OF HAT-P-7
TABLE 1
HJD
RV [m s- 1]
Error [m s- 1]
Spec.a
2454336.73960
2454336.85367
2454337.76212
2454338.77440
2454338.85456
2454339.89886
2454343.83180
2454344.98805
2454983.99020
2454985.02095
2454987.01053
2454988.00940
2455014.95322
2455016.05508
2455042.03696
2455042.98594
2455044.07369
2455000.03550
2455000.07080
2455000.07507
2455000.07933
2455000.08361
2455000.08787
2455000.09214
2455000.09641
2455000.10068
2455013.75919
2455013.76487
2455013.77030
2455013.77573
2455013.78114
2455013.78703
2455013.79245
2455013.79787
2455013.80330
2455013.80871
2455013.81412
2455013.81955
2455013.82496
2455013.83039
2455013.83580
2455013.84121
2455013.84664
2455013.85206
2455013.85748
2455013.86290
2455013.86832
2455013.87374
2455013.88433
2455013.88976
2455013.89518
2455013.90060
2455013.90603
2455013.91145
2455013.92034
2455013.92575
2455013.93118
2455013.93660
2455013.94201
2455013.94744
2455013.95286
2455013.95829
2455013.96372
2455013.96913
2455013.97456
2455013.98000
111.08
58.89
- 236.06
151.06
131.12
- 256.42
- 162.19
84.42
- 77.44
110.41
189.80
- 221.82
- 41.51
4.10
189.91
- 228.74
191.34
196.69
184.21
210.15
195.95
193.11
225.66
210.55
224.15
195.24
74.13
81.76
55.46
44.87
53.93
61.60
42.35
36.87
30.06
25.86
29.68
11.37
5.68
15.44
13.27
3.74
12.79
5.91
- 4.64
0.95
- 10.16
- 3.95
- 7.40
- 1.91
- 22.65
5.42
- 5.14
- 13.89
- 14.44
- 9.34
2.95
- 37.18
- 22.51
- 35.31
- 11.12
5.23
- 59.30
- 63.07
- 55.50
- 64.15
1.72
1.78
1.70
1.54
1.57
1.83
1.75
2.15
2.20
1.96
2.23
2.28
2.08
2.23
2.53
2.40
2.55
12.26
11.29
11.25
11.94
13.30
12.45
11.67
12.23
11.97
8.47
7.53
8.11
7.60
7.71
7.50
6.83
7.31
7.10
6.95
7.44
6.86
6.69
6.63
5.85
5.69
6.91
6.67
6.31
6.07
6.26
5.32
5.70
6.11
5.91
6.32
5.87
6.19
5.90
5.80
5.79
5.53
5.80
5.01
6.17
5.75
5.56
6.04
5.88
5.92
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
1
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
7
Outlandish orbit of HAT-P-7b
TABLE 1 -- Continued
HJD
RV [m s- 1]
Error [m s- 1]
Spec.a
2455013.98542
2455013.99086
2455013.99627
2455014.00486
2455014.01029
2455014.01572
2455014.02114
2455014.02656
2455014.03199
2455014.03742
2455014.04285
2455014.04828
2455014.06427
2455014.06971
2455014.07514
2455014.11226
2455014.11769
2455014.12312
2455014.12855
2455014.13397
- 60.13
- 55.52
- 69.24
- 76.79
- 80.53
- 73.56
- 90.86
- 81.24
- 109.25
- 100.31
- 107.39
- 107.15
- 114.12
- 110.83
- 114.67
- 146.78
- 137.88
- 143.30
- 136.48
- 111.90
5.24
4.96
4.86
5.51
5.47
5.96
5.19
6.44
5.29
5.77
6.12
5.91
5.32
5.39
5.97
5.93
5.84
6.07
5.96
7.02
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
2
NOTE. -- The RV was measured relative to an arbitrary template spectrum specific to each spectrograph; only the differences among the RVs from a single spectrograph are
significant. The uncertainty given in Column 3 is the internal error only and does not account for any possible "stellar jitter."
a (1) Keck/HIRES, (2) Subaru/HDS.
TABLE 2
MODEL PARAMETERS FOR HAT-P-7B
Parameter
Value
Orbital period, P [d]
Midtransit time [HJD]
Transit duration (first to fourth contact) [hr]
Transit ingress or egress duration [hr]
Planet-to-star radius ratio, Rp/R⋆
Orbital inclination, i [deg]
Scaled semimajor axis, a/R⋆
Transit impact parameter
Velocity semiamplitude, K [m s- 1]
Upper limit on eccentricity (99.73% conf.)
e cos ω
e sin ω
Velocity offset, Keck/HIRES [m s- 1]
Velocity offset, Subaru/HDS [m s- 1]
Constant radial acceleration γ [m s- 1 yr- 1]
Projected stellar rotation rate, v sin i⋆ [km s- 1]
Projected spin-orbit angle, λ [deg]
2.2047304 ± 0.0000024
2, 454, 731.67929 ± 0.00043
0.039
- 0.0019 ± 0.0077
0.0037 ± 0.0124
4.006 ± 0.064
0.474+0.061
- 0.093
0.0834+0.0012
- 0.0021
80.8+2.8
- 1.2
3.82+0.39
- 0.16
0.618+0.039
- 0.149
211.8 ± 2.6
- 51.2 ± 3.6
- 4.8 ± 2.5
21.5 ± 2.6
4.9+1.2
- 0.9
182.5 ± 9.4
|
1812.06150 | 2 | 1812 | 2018-12-20T23:11:05 | Synergy between asteroseismology and exoplanet science: an outlook | [
"astro-ph.EP",
"astro-ph.SR"
] | Space-based asteroseismology has been playing an important role in the characterization of exoplanet-host stars and their planetary systems. The future looks even brighter, with space missions such as NASA's TESS and ESA's PLATO ready to take on this legacy. In this contribution, we provide an outlook on the synergy between asteroseismology and exoplanet science, namely, on the prospect of conducting a populational study of giant planets around oscillating evolved stars with the TESS mission. | astro-ph.EP | astro-ph | PHOST "Physics of Oscillating Stars" -- a conference in honour of Prof. H. Shibahashi, 2-7 Sept. 2018, Banyuls-sur-mer (France)
Edited by J. Ballot, S. Vauclair, & G. Vauclair
8
1
0
2
c
e
D
0
2
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
0
5
1
6
0
.
2
1
8
1
:
v
i
X
r
a
Synergy between asteroseismology and exoplanet science:
an outlook
Tiago L. Campante1,2, Susana C. C. Barros1,2, Olivier Demangeon1, Hugo J. da Nóbrega2,3,
James S. Kuszlewicz4,5, Filipe Pereira1, William J. Chaplin6,5, and Daniel Huber7
1 Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, Rua das Estrelas, PT4150-762 Porto, Portugal
Email: [email protected]
2 Departamento de Física e Astronomia, Faculdade de Ciências da Universidade do Porto, Rua do Campo Alegre, s/n, PT4169-007 Porto,
Portugal
3 CAP -- Centre for Applied Photonics, INESC TEC, Porto, Portugal
4 Max Planck Institute for Solar System Research, D-37077 Göttingen, Germany
5 Stellar Astrophysics Centre (SAC), Department of Physics and Astronomy, Aarhus University, Ny Munkegade 120, DK-8000 Aarhus C,
Denmark
6 School of Physics and Astronomy, University of Birmingham, Birmingham B15 2TT, UK
7 Institute for Astronomy, University of Hawai'i, 2680 Woodlawn Drive, Honolulu, HI 96822, USA
Abstract
Space-based asteroseismology has been playing an important role in the characterization of exoplanet-host stars and their
planetary systems. The future looks even brighter, with space missions such as NASA's TESS and ESA's PLATO ready to take
on this legacy. In this contribution, we provide an outlook on the synergy between asteroseismology and exoplanet science,
namely, on the prospect of conducting a populational study of giant planets around oscillating evolved stars with the TESS
mission.
1 Introduction
The asteroseismology revolution initiated by Kepler
(Borucki et al., 2010) is set to continue over the coming
decades with the launches of TESS (Ricker et al., 2015),
PLATO (Rauer et al., 2014), as well as WFIRST (Spergel et al.,
2013), with these missions expected to raise the number of
solar-like oscillators to a few million stars (Huber, 2018).
Note that over 90 % of all detections are expected to be for
evolved stars, with PLATO by far contributing the most de-
tections for dwarfs and subgiants (∼ 80,000). If we combine
this with dedicated ground-based e(cid:29)orts, such as the SONG
network (Grundahl et al., 2017) of 1-meter telescopes, we
are then positive that the synergy between asteroseismology
and exoplanet science can only continue to grow (Campante
et al., 2018).
Synergetic studies of evolved stars are made possible by
even moderate photometric cadences, which can be used to
simultaneously detect transits and stellar oscillations. A very
exciting prospect is that of conducting asteroseismology of
red-giant hosts using the 30-minute cadence of TESS full-
frame images (FFIs). Based on an all-sky stellar and plan-
etary synthetic population (Sullivan et al., 2015), we predict
that solar-like oscillations will be detectable in up to 200 low-
luminosity red-giant branch (LLRGB) stars hosting close-in
giant planets (Campante et al., 2016).
The population of transiting planets around evolved stars
is so far largely unexplored (Huber, 2018). And although
radial-velocity surveys are mostly complete for planets near
or above 1 MJup at > 0.2 AU, there is a dearth of plan-
ets with orbital periods P < 80 d. Nonetheless, Kepler/K2
have discovered several close-in giant planets around LL-
RGB stars (e.g., Grunblatt et al., 2016, 2017), hinting at a
population of warm sub-Jovian planets around evolved stars
that would be accessible to TESS. Kepler/K2 mainly targeted
main-sequence stars, and observed too few LLRGB stars to
detect enough planets for robust statistics. TESS will increase
the number of LLRGB stars with space-based photometry by
one order of magnitude over Kepler/K2, providing an un-
precedented opportunity to address a number of key ques-
tions in exoplanet science, namely:
• The role of stellar (cid:30)ux on hot-Jupiter in(cid:30)ation;
• Giant-planet occurrence as a function of stellar mass
and evolution;
• Correlation between metallicity and giant-planet oc-
currence around evolved stars.
In this work, we start by characterizing the parent popu-
lation of LLRGB stars to be searched for transits based on a
galaxia (Sharma et al., 2011) simulation (Sect. 2). We focus
on the southern ecliptic hemisphere, which will be surveyed
during year 1 of TESS's primary mission. We next imple-
ment a software tool for planetary-transit search based on
a Box-(cid:27)tting Least Squares (BLS) algorithm (Sect. 3). The
tool is tested both for statistical false positive rates and de-
tection sensitivity using arti(cid:27)cial TESS light curves. The for-
mer involves running the code on a su(cid:28)ciently large num-
ber of light curves containing only instrumental/shot noise
and stellar (correlated) signals, namely, granulation and os-
cillations. The latter involves running the code on the same
light curves, although now with injected transits. We con-
clude with a few considerations regarding the use of TESS
photometry alone in candidate vetting (Sect. 4).
2 Parent stellar population
2.1 Synthetic population
We start with an all-sky, magnitude-limited synthetic stel-
lar population generated with galaxia. Output absolute
1
Tiago L. Campante et al.
to an astrophysical false positive. For b < 10◦, the density
of background stars is very high, meaning that any observed
eclipse is more likely to be from a background eclipsing bi-
nary, whereas for b > 20◦ planets should represent a ma-
jority over false positives2 (Sullivan et al., 2015).
Panel (a) of Fig. 3 shows the V -band magnitude distribu-
tion of the parent population (peaking at V ∼ 13 -- 14), while
panel (b) shows a luminosity-color diagram that may be used
to inform target selection. We will likely need to apply a
stricter magnitude cut depending on the actual oscillations
detectability limit. We also assessed the fraction of stars in
the parent population observed over a median of 1 TESS sec-
tor (78 %) and 1 or 2 TESS sectors (93 %). Furthermore, we
note that for 26 % of the stars νmax is above the Nyquist fre-
quency.
3 Automated transit detection:
perfor-
mance assessment
3.1 Arti(cid:27)cial light curves
We generated arti(cid:27)cial light curves for the ∼ 30,000
unique, bona (cid:27)de LLRGB stars in our parent (synthetic) stel-
lar population (see Sect. 2). Generation of the light curves is
performed originally in the frequency domain, after which an
inverse Fourier transform is applied (Kuszlewicz et al., sub-
mitted). We consider only the 30-minute cadence of TESS
FFIs and apply a window function to account for the data
downlink occurring every spacecraft orbit.
We used a photometric noise model for TESS (Sullivan
et al., 2015; Campante et al., 2016) to predict the rms noise
per exposure time. A systematic term of 20 ppm hr1/2 was
included in this calculation. To model the granulation power
spectral density, we adopted a scaled version (to predict TESS
granulation amplitudes) of model F of Kallinger et al. (2014),
which contains two Harvey-like components. No aliased
granulation power was considered. Individual radial, (mixed)
dipole and quadrupole modes were also modeled whenever
νmax < νNyq. Panels (a) and (b) of Fig. 4 respectively dis-
play the power spectral density (PSD) and corresponding
light curve of a V = 10.3 star observed for 27.4 days (or
1 TESS sector). The oscillation bump can be seen around
νmax≈ 167 µHz.
We generated model transit light curves using the Python
package batman3. Assuming circular planetary orbits, we
seeded one planet per star. We next drew orbital periods and
planet radii from uniform distributions spanning the param-
eter space of interest (0.5 to 27.4 days and 4 to 22 R⊕, re-
spectively). Orbital periods were redrawn until no systems
were left within the Roche limit or the stellar envelope. We
assume that all planets transit and draw the impact param-
eter from a uniform distribution de(cid:27)ned over the half-open
interval [0, 1[. Input to batman includes the time of inferior
conjunction, orbital period, planet radius, semi-major axis,
and orbital inclination. A quadratic limb darkening law is
used and its coe(cid:28)cients set to (cid:27)xed values (see Barclay et al.,
2015). We further account for the long integration time by su-
persampling the model 11 times per cadence then integrating
over these subsamples.
2Note that these remarks were made with reference to planet detections
around TESS target stars.
3https://www.cfa.harvard.edu/~lkreidberg/
batman/
Zenodo, 2018
Figure 1: Contamination of RC stars in the initial sample.
magnitudes were converted to apparent magnitudes and ex-
tinction applied. We ended up only retaining stars down to
magnitude 13 in the Johnson -- Cousins IC band. Although
somewhat optimistic, this magnitude cut is used to ensure
that all detectable oscillating LLRGB stars are captured (Cam-
pante et al., 2016). Note that the simulation is undersampled
by a factor of 10 to ease up on the data handling.
We made an initial selection of putative LLRGB stars from
this synthetic population by applying the following cuts on
log g and Teff: 2.7 < log g < 3.5 and Teff < 5500 K. The
upper log g cut ensures that stars are evolved enough to os-
cillate with frequencies detectable with 30-minute-cadence
data (Chaplin et al., 2014). The lower log g cut is purely em-
pirical (Hekker et al., 2011) and leads to contamination of the
sample by red clump (RC) stars (on which more below). This
step was followed by selecting only those stars located in
the southern ecliptic hemisphere. Finally, we determined the
median number of sectors (by considering a number of di(cid:29)er-
ent initial pointings) over which each star would be observed
with TESS using the Python package tvguide1 (each sector
corresponds to a 27.4-day coverage), having discarded stars
that fall o(cid:29) silicon. The (cid:27)nal tally amounts to ∼ 6.8 × 105
stars (after applying the factor-of-10 correction).
We now come back to the issue of the sample contamina-
tion by RC stars. Indeed, 34 % of stars in the above sample are
RC stars (see Fig. 1). We apply an additional cut on the as-
teroseismic observable ∆ν, namely, ∆ν > 10 µHz, in order
to mitigate this contamination e(cid:29)ect. This e(cid:29)ectively leads
to the removal of all RC stars and is the equivalent to raising
the lower log g cut to log g (cid:38) 2.9. We take the resulting sam-
ple of bona (cid:27)de LLRGB stars as our (cid:27)nal, parent (synthetic)
stellar population, which comprises ∼ 3.0 × 105 stars (after
applying the factor-of-10 correction).
2.2 Population characteristics
We now look at the main characteristics of this parent
(synthetic) stellar population. Figure 2 shows the overall stel-
lar radius, mass and metallicity distributions (light red). Dis-
tributions are also shown for a subset of stars lying farther
away from the Galactic plane, i.e., b > 10◦ (light blue; com-
prising 68 % of the parent population). The Galactic latitude
of a target strongly in(cid:30)uences the likelihood of it giving rise
1https://heasarc.gsfc.nasa.gov/docs/tess/
proposal-tools.html#tvguide
2
2.72.82.93.03.13.23.33.43.5log g (dex)010002000300040005000CountAllRCLLRGBPHOST "Physics of Oscillating Stars" conference
(a)
(b)
Figure 3: V -band magnitude distribution (a) and luminosity-color
diagram (b) of the parent stellar population. The luminosity-color
diagram of the parent stellar population (red) is superimposed on
that of the initial sample (gray; see Sect. 2.1).
3.2 BLS algorithm
We search for transits using an updated version of the
pipeline presented in Barros et al. (2016), which makes use of
a Python implementation4 of the BLS algorithm originally
introduced by Kovács et al. (2002). The search is made over
periods ranging from 1 day to 70 % of the light curve dura-
tion and over fractional transit durations ranging from 0.001
to 0.3 with nb = 200 phase bins. Frequency sampling is op-
timized according to δν = 1/(Pmax · nb), where Pmax is the
maximum period searched for. Using the periods and epochs
found by the BLS algorithm, each light curve is phase-folded
and the signal detection e(cid:28)ciency (SDE; Kovács et al., 2002)
computed.
The pipeline searches npass transits per light curve and
sorts them according to the SDE. Results for all the light
curves are also sorted according to the maximum SDE re-
ported for each light curve.
It also tests for the following
features: possibility of a secondary transit/eclipse, sinusoidal
behavior, and mono transit (or an e(cid:29)ective number of tran-
sits that is less than the total number of transits). Provided
4https://github.com/dfm/python-bls
3
(a)
(b)
(c)
Figure 2: Stellar (a) radius, (b) mass and (c) metallicity distributions
of the parent stellar population.
Zenodo, 2018
34567Radius (R_Sun)0500100015002000CountAllb>10º0.81.01.21.41.61.82.02.22.4Mass (M_Sun)0100020003000CountAllb>10º-3-2-10[Fe/H] (dex)01000200030004000CountAllb>10º6789101112131415V (mag)010002000300040005000Count0.50.60.70.80.91.01.11.21.31.41.5J-K0.81.01.21.41.61.82.0log (L/L_Sun)Tiago L. Campante et al.
Table 1: SDE threshold as a function of the statistical false
positive rate.
SDE (All)
SDE (1 sector)
SDE (2 sectors)
Rate
0.1 %
1 %
5 %
7.29
6.34
5.61
6.75
5.98
5.33
7.22
6.39
5.73
(a)
(b)
Figure 4: Power spectral density (a) and corresponding light curve
(b) of a V = 10.3 star observed for 27.4 days (or 1 TESS sector). The
star has νmax ≈ 167µHz. The resolution of the truncated spectrum
(red) is de(cid:27)ned by the duration of the light curve, whereas the orig-
inal spectrum (black) is oversampled. No window function has yet
been applied to the light curve to account for the data downlink.
the (cid:27)tted depth is positive, the pipeline produces a series of
plots for each candidate. Figure 5 shows the pipeline output
for the same arti(cid:27)cial star considered in Fig. 4 (after transit
injection). The two injected transits are correctly recovered.
Another example can be found in Fig. 6, where the pipeline
output is shown for a V = 13.1 star observed for 54.8 days
(or 2 TESS sectors). All three injected transits are correctly
recovered.
3.3 Statistical false positives
We began by running the pipeline on the ∼ 30,000 gen-
erated arti(cid:27)cial light curves (prior to transit injection) in or-
der to assess the rate of statistical false positives. Results are
shown in Fig. 7. From panel (a), we (cid:27)nd that an SDE threshold
of 8.88 produces approximately one statistical false positive
over the ∼ 30,000 light curves or a rate of 0.003 %. Table 1
provides SDE thresholds as a function of the statistical false
positive rate (0.1 %, 1 % and 5 %). Panels (b) and (c) of Fig. 7
emphasize the dependence of the SDE on the sample's lim-
iting magnitude and light curve duration, respectively. The
dependence on the latter is particularly obvious, with shorter
light curves giving rise to lower SDE values (cf. Kovács et al.,
2002). Therefore, Table 1 also provides SDE thresholds when
only 27.4- (1 sector) and 54.8-day-long (2 sectors) light curves
4
are considered.
3.4 Detection sensitivity
We now run the pipeline on the same ∼ 30,000 arti(cid:27)cial
light curves after transit injection in order to assess the de-
tection sensitivity (or survey completeness). Figure 8 shows
the relative di(cid:29)erence between the injected (Pin) and recov-
ered (Pout) orbital periods as a function of the SDE. Vertical
dashed lines mark (from right to left) the SDE thresholds cor-
responding to 0.1 %, 1 % and 5 % statistical false positive rates
when considering 1 TESS sector.
The pipeline sometimes recovers half (0.5 ordinate) or
double (−1 ordinate) the injected period, which tends to hap-
pen close to the aforementioned thresholds. The former sub-
set mostly corresponds to the case of two injected transits
with an orbital period exceeding 70 % of the light curve du-
ration, whereas the latter, less numerous subset mostly cor-
responds to those injected transits with periods shorter than
1 day. These injected periods are out of bounds with respect
to the search parameters of the algorithm. The remaining,
rarer cases seem to be genuine statistical misidenti(cid:27)cations
prompted by the low SDE.
Figure 9 shows the transit detection sensitivity based on
27.4 days of data and a minimum of two transit events, where
we have assumed a statistical false positive rate of 5 %. The
resulting contours demonstrate that even for the most pes-
simistic case of 27.4 days coverage it will be possible to detect
close-in, in(cid:30)ated Jupiters for over 80 % of stars, Jupiter-size
planets for 70 -- 80 % of stars, and large Neptunes only in the
most favorable cases.
3.4.1 Out-of-transit (cid:30)ux modulation
We test for the presence of sinusoidal behavior in the light
curve by (cid:27)rst (cid:27)tting a sine function to the binned, folded light
curve (on the best period). We then perform a linear regres-
sion (after linearization) to obtain the coe(cid:28)cient of determi-
nation, r2. Here, we apply this procedure to an arti(cid:27)cial sys-
tem resembling the con(cid:27)rmed Kepler-91 planetary system,
for which ellipsoidal variations have been measured that are
caused by a close-in giant planet (Lillo-Box et al., 2014; Bar-
clay et al., 2015). We injected a sine function with period
Pin/2 and phase 0 at minimum into the light curve (Pfahl
et al., 2008), having varied its amplitude. The star is observed
for 27.4 days only (or 1 TESS sector). A notional amplitude
of 100 ppm of the ellipsoidal modulation (see Fig. 10) leads
to r2 = 0.22, above the r2 = 0.1 threshold adopted within
the pipeline to (cid:30)ag potential sinusoidal behavior. This am-
plitude is slightly higher than the approximate upper limit
of 75 ppm measured for Kepler-91. We conclude by noting
that it is usually necessary to remove instrumental trends and
rotational modulation due to spots/plages prior to any tran-
sit search. Such detrending of the light curve may (depend-
ing on the period) also remove part of the out-of-transit (cid:30)ux
modulation, thus making detection of ellipsoidal variations
more challenging.
Zenodo, 2018
PHOST "Physics of Oscillating Stars" conference
Figure 5: Pipeline output for the same arti(cid:27)cial star considered in Fig. 4. The light curve is shown in the top left panel with both (correctly)
recovered transits in red. Notice the gap at ≈ 13.7 d due to the data downlink. The BLS periodogram is shown in the top right panel with the
vertical dashed line indicating the best period, as determined by the algorithm. The bottom left panel displays the phase-folded light curve
using the best period (blue) and a binned version of it (red). The bottom middle and right panels simply zoom in on the phase-folded light
curve at the locations of the primary and possible secondary, respectively.
Figure 6: Pipeline output for a V = 13.1 star observed for 54.8 days (or 2 TESS sectors). Panels are the same as in Fig. 5.
3.4.2
Secondary transit/eclipse
We test for the presence of a secondary transit/eclipse
by (i) looking for two closely spaced periods (at most 0.5
days apart) and (ii) assessing whether the corresponding
depths di(cid:29)er by at least 10 % (relative to the larger of the two
depths). The former step also involves checking if a newly
detected period matches the 1st overtone of the primary (i.e.,
/2). Figure 11 illustrates this procedure. The light
P primary
out
Zenodo, 2018
curve has nine injected primary transits. These are easily re-
covered during the (cid:27)rst period search despite the clear pres-
ence of a secondary at phase 0.5 (see top panel). A planet-to-
star (cid:30)ux ratio of 5× 10−4 was assumed when simulating the
secondary. Such an arbitrarily large planet-to-star (cid:30)ux ratio
-- as a term of comparison, the occultation depth of Kepler-
91b is ∼ 50 ppm -- is employed here for illustrative purposes
only and no attention has been paid to determine whether or
5
0.20.00.20.40.60.8Phase0.99850.99900.99951.00001.0005Relative fluxPhase folding for trial period-0.10-0.050.000.050.10Phase0.99850.99900.99951.00001.0005Zooming in for primary search05101520Period [days]0.00.20.40.60.81.01.21.4BLS power1e4Periodogram0.400.450.500.550.60Phase0.00050.00100.00150.0020+9.99e1Zooming in for secondary search051015202530Time [days]0.99850.99900.99951.00001.0005Relative fluxLight curvePeriod: 16.89565 SDE: 6.26246Duration: 0.75211 Depth: 0.00067Epoch: 0.33791 Rp/Rs: 0.02579STAR art_231980 MAG 10.2924 PER. SEARCH NO. 10.20.00.20.40.60.8Phase0.9970.9980.9991.0001.0011.002Relative fluxPhase folding for trial period-0.08-0.06-0.04-0.020.000.020.040.060.08Phase0.9970.9980.9991.0001.0011.002Zooming in for primary search0510152025303540Period [days]0.20.40.60.81.01.21.41.61.82.0BLS power1e4Periodogram0.400.450.500.550.60Phase0.0010.0020.0030.0040.005+9.97e1Zooming in for secondary search0102030405060Time [days]0.9970.9980.9991.0001.0011.002Relative fluxLight curvePeriod: 19.30344 SDE: 9.19455Duration: 0.73683 Depth: 0.00098Epoch: 12.54723 Rp/Rs: 0.03129STAR art_2514657 MAG 13.1009 PER. SEARCH NO. 1Tiago L. Campante et al.
Figure 8: Relative di(cid:29)erence between injected (Pin) and recovered
(Pout) orbital periods as a function of the SDE. Only systems with
two or more detected transits and measured positive depths have
been displayed. Vertical dashed lines mark (from right to left) the
SDE thresholds corresponding to 0.1 %, 1 % and 5 % statistical false
positive rates when considering 1 TESS sector (see Table 1). Systems
observed over 1 (gray) or more (green) TESS sectors are identi(cid:27)ed.
Figure 9: Transit detection sensitivity based on 27.4 days of data
and a minimum of two transit events. We assume a statistical false
positive rate of 5 % (or an SDE threshold of 5.33; see Table 1).
not it is physically sound. When performing the second pe-
riod search (after masking out all primary transits from the
light curve), we notice how the BLS power associated with
the true period is brought down relative to that of its 1st over-
tone (see bottom panel). Detection of the 1st overtone of the
primary (after its removal) is a telltale sign of the presence of
a secondary transit/eclipse.
4 Candidate ve(cid:29)ing
We will be using TESS photometry alone in a (cid:27)rst attempt
to separate the expected transit-like signals from (the notion-
ally many) astrophysical false positives (due, e.g., to EBs and
BEBs) and systematic false alarms. A clear distinction will
not always be possible, of course, but this will aid retaining
those cases we think are due to genuine transits of short-
Zenodo, 2018
(a)
(b)
(c)
Figure 7: Statistical false positive rates. Panel (a): Number of sta-
tistical false positives as a function of the SDE. Horizontal dashed
lines indicate 1, 10, and 100 false positives. Panels (b) and (c): Nor-
malized histograms of the SDE depicting its dependence on limiting
magnitude and light curve duration, respectively.
6
24681012140.40.60.81.01.21.41.62468101214Orbital Period (Days)0.40.60.81.01.21.41.6Planet Radius (Jupiter Radii)0.20.30.40.50.60.70.80.924681012140.40.60.81.01.21.41.60.00.30.61.0Fraction of SDE>5.33 DetectionsPHOST "Physics of Oscillating Stars" conference
Acknowledgments
The project leading to this publication has received fund-
ing from the European Union's Horizon 2020 research and
innovation programme under the Marie Skłodowska-Curie
grant agreement No. 792848 (PULSATION). This work was
supported by FCT -- Fundação para a Ciência e a Tecnolo-
gia through national funds and by FEDER through COM-
PETE2020 -- Programa Operacional Competitividade e Inter-
nacionalização by these grants: UID/FIS/04434/2013 & POCI-
01-0145-FEDER-007672 and PTDC/FIS-AST/28953/2017 &
POCI-01-0145-FEDER-028953. S.C.C.B. acknowledges Inves-
tigador FCT contract IF/01312/2014/CP1215/CT0004. The re-
search leading to the presented results has received funding
from the European Research Council under the European
Community's Seventh Framework Programme (FP7/2007-
2013) / ERC grant agreement no. 338251 (StellarAges).
References
Barclay, T., Endl, M., Huber, D., Foreman-Mackey, D.,
Grundahl,
F.,
2010, Science, 327, 977.
Cochran, W. D., et al. 2015, ApJ, 800, 46.
594, A100.
Grunblatt, S. K., Huber, D., Gaidos, E., Lopez, E. D., Howard,
Chaplin, W. J., Elsworth, Y., Davies, G. R., Campante, T. L.,
Campante, T. L., Scho(cid:27)eld, M., Kuszlewicz, J. S., Bouma, L.,
Barros, S. C. C., Demangeon, O., & Deleuil, M. 2016, A&A,
Borucki, W. J., Koch, D., Basri, G., Batalha, N., Brown, T., et al.
Campante, T. L., Santos, N. C., & Monteiro, M. J. P. F. G. 2018,
Asteroseismology and Exoplanets: Listening to the Stars
and Searching for New Worlds, 49.
Chaplin, W. J., et al. 2016, ApJ, 830, 138.
Handberg, R., et al. 2014, MNRAS, 445, 946.
A. W., et al. 2017, AJ, 154, 254.
B. J., et al. 2016, AJ, 152, 185.
Fredslund Andersen, M., Christensen-
Dalsgaard, J., Antoci, V., Kjeldsen, H., et al. 2017, ApJ, 836,
142.
Hekker, S., Gilliland, R. L., Elsworth, Y., Chaplin, W. J., De
Ridder, J., et al. 2011, MNRAS, 414, 2594.
to the Stars and Searching for New Worlds, 49, 119.
et al. 2014, A&A, 570, A41.
Huber, D. 2018, Asteroseismology and Exoplanets: Listening
Grunblatt, S. K., Huber, D., Gaidos, E. J., Lopez, E. D., Fulton,
Kallinger, T., De Ridder, J., Hekker, S., Mathur, S., Mosser, B.,
Kovács, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369.
Lillo-Box, J., Barrado, D., Moya, A., Montesinos, B., Montal-
bán, J., et al. 2014, A&A, 562, A109.
Pfahl, E., Arras, P., & Paxton, B. 2008, ApJ, 679, 783.
Rauer, H., Catala, C., Aerts, C., Appourchaux, T., Benz, W.,
et al. 2014, Experimental Astronomy, 38, 249.
Ricker, G. R., Winn, J. N., Vanderspek, R., Latham, D. W.,
Bakos, G. Á., et al. 2015, Journal of Astronomical Tele-
scopes, Instruments, and Systems, 1, 014003.
Sharma, S., Bland-Hawthorn, J., Johnston, K. V., & Binney, J.
Spergel, D., Gehrels, N., Breckinridge, J., Donahue, M.,
2011, ApJ, 730, 3.
Dressler, A., et al. 2013, ArXiv e-prints.
neau, D., Deming, D., et al. 2015, ApJ, 809, 77.
Sullivan, P. W., Winn, J. N., Berta-Thompson, Z. K., Charbon-
7
(a)
(b)
Figure 10: Light curve (a) and phase-folded light curve (b) of an
arti(cid:27)cial system resembling the Kepler-91 system. The star is ob-
served for 27.4 days (or 1 TESS sector). A sine function with pe-
riod Pin/2 and phase 0 at minimum has been injected into the light
curve. Its amplitude was set to 100 ppm. An ellipsoidal modulation
is clearly seen in the right panel.
period gas giants around LLRGB stars.
Housekeeping operations such as retaining only those
systems with two or more detected transits as well as mea-
sured positive depths will be implemented. An SDE threshold
will be adopted, its value depending on the number of TESS
sectors (cf. Sect. 3.3). Any measured transit depth in excess
of 1 % will be attributed to an eclipsing binary (a 2 RJ planet
transiting an LLRGB will at most cause a ∼0.5 % (cid:30)ux reduc-
tion). As illustrated in the previous section, the pipeline tests
for a number of additional features, namely, the presence of a
secondary transit/eclipse and sinusoidal behavior. Although
not decisive with respect to the vetting procedure, these (cid:30)ags
provide useful information that can be used during the de-
tailed transit (cid:27)tting. We will pay particular attention, at low
SDE, to possible period misidenti(cid:27)cations (due to an out-of-
bounds true period or else statistical in nature) by carefully
inspecting the light curve and associated BLS periodogram.
We also advocate for the independent (cid:27)tting -- as part of the
detailed transit (cid:27)tting -- of the phase-folded odd and even
primary transits, since a signi(cid:27)cant di(cid:29)erence in their depths
may indicate we are dealing with an eclipsing binary.
Zenodo, 2018
Tiago L. Campante et al.
(a)
(b)
Figure 11: Pipeline output for a V = 13.0 star observed for 109.6 days (or 4 TESS sectors). A primary (at phase 0) and a secondary (at phase
0.5) were injected into the light curve. Panels refer to the (cid:27)rst (a) and second (b) period search. Notice that all primary transits have been
masked out from the light curve prior to the second period search.
8
Zenodo, 2018
|
1802.07725 | 2 | 1802 | 2018-04-09T15:12:09 | Increased Heat Transport in Ultra-Hot Jupiter Atmospheres Through H$_2$ Dissociation/Recombination | [
"astro-ph.EP"
] | A new class of exoplanets is beginning to emerge: planets whose dayside atmospheres more closely resemble stellar atmospheres as most of their molecular constituents dissociate. The effects of the dissociation of these species will be varied and must be carefully accounted for. Here we take the first steps towards understanding the consequences of dissociation and recombination of molecular hydrogen (H$_2$) on atmospheric heat recirculation. Using a simple energy balance model with eastward winds, we demonstrate that H$_2$ dissociation/recombination can significantly increase the day$-$night heat transport on ultra-hot Jupiters (UHJs): gas giant exoplanets where significant H$_2$ dissociation occurs. The atomic hydrogen from the highly irradiated daysides of UHJs will transport some of the energy deposited on the dayside towards the nightside of the planet where the H atoms recombine into H$_2$; this mechanism bears similarities to latent heat. Given a fixed wind speed, this will act to increase the heat recirculation efficiency; alternatively, a measured heat recirculation efficiency will require slower wind speeds after accounting for H$_2$ dissociation/recombination. | astro-ph.EP | astro-ph | Draft version April 10, 2018
Typeset using LATEX twocolumn style in AASTeX62
8
1
0
2
r
p
A
9
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
5
2
7
7
0
.
2
0
8
1
:
v
i
X
r
a
Increased Heat Transport in Ultra-Hot Jupiter Atmospheres Through
H2 Dissociation/Recombination
Taylor J. Bell1, ∗ and Nicolas B. Cowan1, 2, ∗
1Department of Physics, McGill University, 3600 rue University, Montr´eal, QC H3A 2T8, Canada
2Department of Earth & Planetary Sciences, McGill University, 3450 rue University, Montr´eal, QC H3A 0E8, Canada
(Received 2018 January 30; Revised 2018 April 04; Accepted 2018 April 07)
Submitted to ApJL
ABSTRACT
A new class of exoplanets is beginning to emerge: planets whose dayside atmospheres more closely
resemble stellar atmospheres as most of their molecular constituents dissociate. The effects of the
dissociation of these species will be varied and must be carefully accounted for. Here we take the
first steps towards understanding the consequences of dissociation and recombination of molecular
hydrogen (H2) on atmospheric heat recirculation. Using a simple energy balance model with eastward
winds, we demonstrate that H2 dissociation/recombination can significantly increase the day–night
heat transport on ultra-hot Jupiters (UHJs): gas giant exoplanets where significant H2 dissociation
occurs. The atomic hydrogen from the highly irradiated daysides of UHJs will transport some of the
energy deposited on the dayside towards the nightside of the planet where the H atoms recombine
into H2; this mechanism bears similarities to latent heat. Given a fixed wind speed, this will act to
increase the heat recirculation efficiency; alternatively, a measured heat recirculation efficiency will
require slower wind speeds after accounting for H2 dissociation/recombination.
Keywords: planets and satellites: atmospheres - planets and satellites: gaseous planets - methods:
analytical - methods: numerical
1. INTRODUCTION
Most gas giant exoplanets have atmospheres domi-
nated by molecular hydrogen (H2). However, on planets
where the temperature is sufficiently high, a significant
fraction of the H2 will thermally dissociate; one may call
these planets ultra-hot Jupiters (UHJs). Only a handful
of known planets have dayside temperatures this high,
but the TESS mission is expected to discover hundreds
more as it includes many early-type stars (George Zhou,
private communication 2017). These UHJs are an in-
teresting intermediate between stars and cooler planets,
and they will allow for useful tests of atmospheric mod-
els.
Corresponding author: Taylor J. Bell
[email protected]
∗ McGill Space Institute; Institute for Research on Exoplanets;
Centre for Research in Astrophysics of Quebec
At these star-like temperatures, the H− bound-free
and free-free opacities should play an important role in
the continuum atmospheric opacity which has recently
been detected in dayside secondary eclipse spectra (Bell
et al. 2017; Arcangeli et al. 2018). These recently re-
ported detections of H− opacity provide evidence that
H2 is dissociating in the atmospheres of gas giants at
this temperature range. However, the thermodynamical
effects of H2 dissociation/recombination have yet to be
explored.
Both theoretically (e.g. Perez-Becker & Showman
2013; Komacek & Showman 2016) and empirically (e.g.
Zhang et al. 2018; Schwartz et al. 2017), we expect
the day–night temperature contrast on hot Jupiters
to increase with increasing stellar irradiation; temper-
ature gradients (cid:38) 1000 K can be expected for UHJs.
As temperatures vary drastically between day and
night, the local thermal equilibrium (LTE) H2 disso-
ciation fraction will also vary. The recombination of H
into H2 is a remarkably exothermic process, releasing
2
Bell & Cowan
"Top-Down" View
advanced circulation models to explore this problem in
a more rigorous and quantitative manner.
2. ENERGY TRANSPORT MODEL
2.1. Heating Terms
First, let be the energy per unit area of a parcel
of gas. Ignoring H2 dissociation/recombination and any
internal heat sources, and assuming the gas parcel cools
radiatively, energy conservation gives
Figure 1. A cartoon showing a "top-down" view of the
expected dissociation and recombination of H2 on an ultra-
hot Jupiter (UHJ). The orbital direction and the direction
of winds on the planet are indicated with black arrows.
q = 2.14×108 J kg−1 (Dean 1999); this is 100× more po-
tent than the latent heat of condensation for water. For
reference, latent heat is responsible for approximately
half of the heat recirculation on Earth (L/(cp∆T ) ∼ 1),
while the effect of H2 dissociation/recombination should
be even stronger for UHJs (q/(cp∆T ) ∼ 102).
Building on this intuition, we might expect that H will
recombine into H2 as gas carried by winds flows eastward
from the sub-stellar point, significantly heating the east-
ern hemisphere of the planet. As the gas continues to
flow around to the dayside, the H2 will again dissoci-
ate and significantly cool the western hemisphere. A
cartoon depicting this layout is shown in Figure 1.
If
unaccounted for while modelling a phasecurve, this may
manifest itself as an "unphysically" large eastward offset
as was previously reported for WASP-12b (Cowan et al.
2012).
A large number of circulation models have been devel-
oped for studying exoplanet atmospheres, ranging from
simple energy balance models (e.g. Cowan & Agol 2011)
to more advanced general circulation models (e.g. Show-
man et al. 2009; Rauscher & Menou 2010; Amundsen
et al. 2014; Zhang & Showman 2017; Heng & Kitzmann
2017; Dobbs-Dixon & Cowan 2017). To our knowledge,
however, no published general circulation models ac-
count for the cooling/heating due to the energies of H2
dissociation/recombination (although some planet for-
mation models do account for this, e.g. Berardo et al.
2017). Here we aim to qualitatively explore the effects
of H2 dissociation/recombination using a simple energy
balance model adapted from that described by Cowan &
Agol (2011), using code based on that implemented by
Schwartz et al. (2017). We leave it to those with more
d
dt
= Fin − Fout,
with Fin and Fout given by
Fin = (1 − AB)F∗ sin θ max
Fout = σT 4.
(cid:16)
(cid:17)
cos Φ(t), 0
,
The planet's Bond albedo is given by AB, θ is the co-
latitude of the gas parcel, T is the temperature of the
gas parcel, and σ is the Stefan-Boltzmann constant. The
incoming stellar flux is given by F∗ = σT 4∗,eff (R∗/a)2,
where T∗,eff is the stellar effective temperature, R∗ is the
stellar radius, and a is the planet's semi-major axis. The
stellar hour angle, Φ(t), incorporates both advection and
planetary rotation.
In order to include H2 dissociation/recombination, we
add a new term accounting for the energy flux from these
effects. This can be done with
d
dt
= Fin − Fout −
dQ
dt
,
(1)
where the energy per unit area stored by H2 dissociation
is given by
Q = qχHΣ,
where Σ is the mass per unit area of H and H2 in the
parcel of gas (in kg m−2), q = 2.14×108 J kg−1 is the H2
bond dissociation energy per unit mass at 0 K, and χH
is the dissociation fraction of the gas. χH = 1 means the
gas is completely dissociated (all atomic). Assuming the
gas parcel is in hydrostatic equilibrium, we can use
(cid:90) ∞
Σ =
z0
ρ(z)dz = (P0/g)
where z0 is some reference height, P0 is the atmospheric
pressure corresponding to z0, and ρ is the density of the
gas. This then allows us to rewrite Q as
Q = (P0/g)qχH.
The time derivative of Q is then
dQ
dt
= (P0/g)q
dχH
dt
= (P0/g)q
(cid:12)(cid:12)(cid:12)(cid:12)T
dχH
dT
dT
dt
,
(2)
OrbitalDirectionHH2DissociationRecombinationH2 Dissociation/Recombination in UHJs
3
where we have assumed the gas parcel's P0/g remains
constant, and where we have made use of the chain rule
to expand dχH/dt.
We model the LTE H2 dissociation fraction by solving
the Saha equation as stated in Appendix A of Berardo
et al. (2017) for χH, assuming the atmosphere consists
of only H and H2:
χH(P, T ) =
nH
nH + nH2
=
2
1 + √1 + 4Y
,
(3)
where nH and nH2 are the number densities of H and
H2,
Y =
T −3/2 exp(q/2mHkBT )P
B h−3Θrot
2(πmH)3/2k5/2
,
where mH is the mass of the hydrogen atom, kB is
the Boltzmann constant, h is the Planck constant,
Θrot = 85.4 K is rotational temperature of H2 (Hill
1986), P is the gas pressure, and T is the tempera-
ture of the gas (in K). The LTE dissociation fraction is
plotted in the top panel of Figure 2.
We can then find dχH/dT using the chain rule:
dχH
dT
=
dχH
dY
dY
dT
.
After some simplification, we then determine
(cid:18) 3
2
χ2
HY
dχH
dT
=
To a good degree of accuracy, Equations (3) and (4) can
be approximated at P = 0.1 bar using
T −1 + (q/2mHkB)T −2
√1 + 4Y
(cid:32)
1 + erf
1
2
(cid:19)(cid:33)
(cid:18) T − µ
σ√2
(cid:19)
.
(4)
(5)
(6)
χH(0.1 bar, T ) =
and
dχH(0.1 bar, T )
dT
=
1
σ√2π
e−(T−µ)2/2σ2
where σ = 471 K and µ = 3318 K, and erf is the error
function; this approximation offers a 70% increase in
computation speed. It should be noted that we assume
that this H2 dissociation/recombination occurs instan-
taneously since the timescale in the temperature regime
of UHJs at 0.1 bar is ∼10−3 s (Rink 1962; Shui 1973).
2.2. Thermal Energy
We assume that the planet's energy is stored entirely
as thermal energy, as is done in other simple energy
balance models (e.g. Cowan & Agol 2011; Pierrehumbert
2010). This assumption means
(cid:16)
d
dt
(cid:17)
(cpT Σ) =
(P0/g)cpT
d
dt
=
d
dt
(cid:18)
dT
dt
(cid:18)
= (P0/g)
cp
+ T
= (P0/g)
dT
dt
cp + T
(cid:19)
(cid:12)(cid:12)(cid:12)(cid:12)T
dcp
dt
dcp
dT
(cid:19)
.
(7)
where cp is the specific heat capacity of the gas, we have
again assumed the gas parcel's P0/g remains constant,
and we have used the chain rule to expand dcp/dt.
The specific heat capacity of the gas will change as a
function of temperature due to the slightly different val-
ues for H and H2 as well as the variations in the specific
heat capacity of H2 as a function of temperature (Chase
1998); any model properly accounting for H2 dissoci-
ation should account for this effect. In our model, we
assume the atmosphere is made entirely of hydrogen and
model the specific heat capacity of the gas by assuming
it is well mixed so that
cp = cp,HχH + cp,H2(1 − χH),
where both χH and cp,H2 are functions of temperature.
The temperature derivative of cp is then given by
(cid:12)(cid:12)(cid:12)(cid:12)T
dcp
dT
= (cp,H − cp,H2 )
dχH
dT
2.3. Putting Everything Together
Putting together Equations (1), (2) and (7), we get
Fin − Fout − (P0/g)
= (P0/g)
dT
dt
dT
dt
q
dχH
dT
cp + T
dcp
dT
(cid:19)
.
(cid:12)(cid:12)(cid:12)(cid:12)T
After solving for dT /dt, we find
dT
dt
= (Fin− Fout)(P0/g)
−1
cp + T
dcp
dT
(cid:18)
Finally, a gas cell can then be updated using
(cid:18)
∆T =
(P0/g)
∆t(Fin − Fout)
cp + T
+ q
dcp
dT
dχH
dT
(cid:12)(cid:12)(cid:12)(cid:12)T
(cid:19)−1
.
(cid:12)(cid:12)(cid:12)(cid:12)T
+ q
dχH
dT
(cid:19) .
(8)
(cid:12)(cid:12)(cid:12)(cid:12)T
Note that the entire sum in the denominator can instead
be thought of as the specific heat capacity of a gas com-
prised of a mixture of H and H2 in thermal equilibrium.
The relative importance of the terms in this sum are
shown in the bottom two panels of Figure 2.
(cid:18)
(cid:18)
(cid:12)(cid:12)(cid:12)(cid:12)T
.
(cid:12)(cid:12)(cid:12)(cid:12)T
(cid:12)(cid:12)(cid:12)(cid:12)T
(cid:19)
4
Bell & Cowan
T∗,eff = 6360 K, R∗ = 1.657 R(cid:12), P = 1.0914203 days
(Collins et al. 2017) and AB = 0.27 (Schwartz et al.
2017). We have also assumed a photospheric pressure
of 0.1 bar, the approximate pressure probed by NIR ob-
servations of WASP-12b (Stevenson et al. 2014), which
gives a radiative timescale of a few hours (similar to
the observed timescales for eccentric hot Jupiters, e.g.
Lewis et al. 2013; de Wit et al. 2016). Wind speeds for
WASP-12b have not been directly measured, but typical
values for hot Jupiters are on the order of 1 km s−1 (e.g.
Koll & Komacek 2017); for that reason, we focus on
wind speeds around this order of magnitude.
First,
let's explore the effects of H2 dissociation/
recombination at a spatially resolved scale. Figure 3
shows temperature and H2 dissociation maps for three
different wind speeds.
In the limit of infinite wind
speeds, there will be no temperature gradients and H2
dissociation/recombination will not play a role. In the
limit of an atmosphere in radiative equilibrium (wind
speed = 0), there will be no variation in the tempera-
ture of a parcel and H2 dissociation/recombination will
play no role. Outside of these two unphysical limits,
H2 dissociation/recombination will always be occurring
somewhere on UHJs.
We now consider phasecurve observations - this re-
quires that we convolve the planet map with a visibil-
ity kernel at each orbital phase (Cowan et al. 2013),
which acts as a low-pass filter. Figure 4 shows model
phasecurves for three wind speed; this figure shows
that H2 dissociation/recombination can have a signifi-
cant effect. At a constant wind speed, the first obvi-
ously affected observable when accounting for dissoci-
ation/recombination is the increased offset of the peak
in the phasecurve towards the east (the same direction
as the prescribed wind). Another affected observable
is the amplitude of the phase variations which is re-
duced when H2 dissociation/recombination is included.
Also, a Fourier decomposition shows that nearly all of
the power in the phasecurves accounting for H2 disso-
ciation/recombination is in the first and second order
Fourier series terms (1forb and 2forb). Finally, Figure 5
shows the trends in phase offset and nightside temper-
ature for two wind speeds, both accounting for and ne-
glecting H2 dissociation/recombination.
4. MODEL ASSUMPTIONS
With simplistic models, many important effects are
necessarily swept under the rug. Here we aim to lift
up the rug and shine a light on our assumptions to
aid future work. While many of these assumptions
will change the quantitative effects of H2 dissociation/
recombination, we expect that the overall qualitative im-
Figure 2. Top: The LTE dissociation fraction of H2 in a
parcel of gas. Middle: A demonstration of the relative im-
portance of the H2 dissociation/recombination term in Equa-
tion (8). For 2300 (cid:46) T (cid:46) 4300, the energy absorbed by H2
dissociation is greater than the energy stored as heat. Typ-
ical hot Jupiters are too cool to be affected by H2 dissocia-
tion/recombination, but these processes should dominate on
UHJs. Bottom: An inset showing the specific heat capac-
ity of a gas composed of H and H2 in LTE (the same black
line from the middle panel), the specific heat capacities of H
and H2 where they are able to exist in equilibrium, and the
additional T (dcp/dT ) term. All panels assume a pressure of
0.1 bar.
3. SIMULATED OBSERVATIONS AND
QUALITATIVE TRENDS
We now explore the effects of this new term in the
differential equation governing the temperature of a gas
cell. For this purpose, we create a latitude+longitude
HEALPix grid where each parcel's temperature is up-
dated using Equation (8) with code based on that de-
veloped by Schwartz et al. (2017).
While Cowan & Agol (2011) were able to explore
their model using dimensionless quantities, our up-
dated model requires that we use dimensioned variables.
We therefore adopt the values of the first discovered
UHJ, WASP-12b (Hebb et al. 2009).
In particular,
we set Rp = 1.90 RJ , a = 0.0234 AU, Mp = 1.470 MJ ,
0.000.250.500.751.00χ05101520104JK−1kg−1H2Diss.cpTdcpdT100020003000400050006000T(K)012104JK−1kg−1cpcp,H2cp,HTdcpdTH2 Dissociation/Recombination in UHJs
5
0.1 km s−1
Equatorial Zonal Wind Speeds
1.0 km s−1
10 km s−1
Temperature
Maps
H2 Dissociation
Maps
Figure 3. Planetary maps, showing temperature and H2 dissociation fraction, assuming different eastward zonal wind speeds.
The dayside hemisphere is shown on the left side of each map with north at the top.
Figure 4. Model bolometric phasecurves assuming different
eastward zonal wind velocities, ignoring eclipses and tran-
sits. The thick, red lines show the expected phasecurve ac-
counting for H2 dissociation/recombination, while the thin-
ner, black models neglect these processes. Secondary eclipse
would occur at a phase of 0.0, while a transit would occur
at 0.5.
(cid:112)R∗/a), given theoretical bolomet-
Figure 5. A figure showing the trends in nightside appar-
ent temperature and phase offset as a function of irradiation
temperature (T0 ≡ T∗,eff
ric phasecurve measurements. Thick, red lines show mod-
els including H2 dissociation/recombination for WASP-12b,
while thin, black lines show models neglecting these effects.
Models sharing the same wind speed share linestyles, and all
models assume a Bond albedo of 0.3 (which is typical for hot
Jupiters; Zhang et al. 2018; Schwartz et al. 2017). A vertical
dotted line shows the location of WASP-12b.
500150025003500T(K)500150025003500T(K)500150025003500T(K)0.00.51.0χ0.00.51.0χ0.00.51.0χ0200400600800Fp/F∗(ppm)0.1kms−10200400600800Fp/F∗(ppm)1.0kms−10.00.20.40.60.81.0OrbitalPhase0200400600800Fp/F∗(ppm)10kms−1050010001500200025003000Tnight(K)vwind=10kms−1vwind=1kms−120003000400050006000T0(K)0◦10◦20◦30◦40◦50◦60◦BolometricPhaseOffset6
Bell & Cowan
pact of increased heat recirculation will be robust to
these assumptions.
One important piece of physics that we have ignored
(beyond a simple assumption of a 0.1 bar photosphere) is
atmospheric opacity. As Dobbs-Dixon & Cowan (2017)
demonstrated, variations in opacity sources as a function
of longitude can change the depth of the photosphere by
an order of magnitude or more. Changing the H2 dis-
sociation fraction will change the importance of H− as
an opacity source, and other standard opacity sources
(e.g. H2O and CO) will also likely be important, es-
pecially towards the cooler nightside. The insignificant
detection of H2O on the dayside of WASP-12b (Steven-
son et al. 2014) but significant detection in the planet's
transmission spectrum (Kreidberg et al. 2015) clearly
demonstrates that opacity sources should be expected
to change on UHJs. Several of the standard molecular
opacity sources will also overlap with the far broader
H− absorption, which complicates a definitive detec-
tion of H− using broadband photometry, such as with
Spitzer /IRAC. The formation of clouds on the nightside
of the planet would further complicate the interpreta-
tion of observed phasecurves, increasing the albedo of
the west terminator while also insulating the nightside.
While we have accounted for variations in the radiative
timescale as a function of temperature, we have not ac-
counted for changes due to varying opacity sources.
Additionally, as we have assumed all photons are emit-
ted at a 0.1 bar photosphere, the effects of the atmo-
sphere's T-P profile have been neglected. As the H2
dissociation fraction has a fairly weak dependence on
gas pressure, the bulk of vertical variations in the H2
dissociation fraction will likely be controlled by the ver-
tical temperature gradient. Due to the lower density of
the dissociated gas, one may expect vertical advection
on UHJs where temperature decreases with altitude. In-
terestingly, however, observations of most UHJs are best
explained by atmospheres with thermal inversions (Ar-
cangeli et al. 2018; Evans et al. 2017) or at least approx-
imately isothermal profiles on the dayside (Crossfield
et al. 2012; Cowan et al. 2012). Any non-isothermal T-P
profile will alter the specifics of how efficiently heat is re-
distributed across the planet as different layers in a gas
column will dissociate/recombine at different locations.
Also, as we have neglected atmospheric opacity, we have
assumed that each gas parcel emits as a blackbody with
a single temperature.
Further, due to the changing scale height of the at-
mosphere at different latitudes and longitudes due to
changes in temperature and H2 dissociation fraction,
there will likely be a tendency for gas to flow away from
the sub-stellar point both zonally and meridionally. This
is not accounted for in our toy model and would re-
quire a general circulation model. Instead, we have cho-
sen eastward winds as they are predicted, and seen, for
most hot Jupiters (e.g. Showman & Guillot 2002; Zhang
et al. 2018), although there are some exceptions (e.g.
Dang et al. 2018). Similarly, our assumption of solid-
body atmospheric rotation is clearly an over simplifica-
tion which will need to be addressed in future work. Our
model is also unable to predict the wind speeds of UHJs
which would require the implementation of various drag
sources, such as magnetic drag which Menou (2012) sug-
gests will dominate at these high temperatures.
Also, we have assumed all heating is due to H2 dis-
sociation and radiation from the host star, neglecting
other heat sources such as residual heat from formation
(which should be negligible for planets older than 1 Gyr;
Burrows et al. 2006) as well as tidal, viscous, and ohmic
heating. We have also neglected the presence of helium
which will partially dilute the strength of H2 dissoci-
ation/recombination as only ∼80% of the atmosphere
will be hydrogen. Finally, we have assumed that the
planet has a uniform albedo which will not be the case
in general (e.g. Esteves et al. 2013; Demory et al. 2013;
Angerhausen et al. 2015; Parmentier et al. 2016).
5. DISCUSSION AND CONCLUSIONS
A new class of exoplanets is beginning to emerge:
planets whose dayside atmospheres resemble stellar at-
mospheres as their molecular constituents thermally dis-
sociate. The impacts of this dissociation will be var-
ied and must be carefully accounted for. Here we have
shown that the dynamical dissociation and recombina-
tion of H2 will play an important role in the heat re-
circulation of ultra-hot Jupiters.
In the atmospheres
of ultra-hot Jupiters, significant H2 dissociation occurs
on the highly irradiated dayside, absorbing some of the
incident stellar energy and transporting it towards the
nightside of the planet where the gas recombines. Given
a fixed wind speed, this will act to increase the heat
recirculation efficiency; alternatively, a measured heat
recirculation efficiency will require slower wind speeds
once H2 dissociation/recombination has been accounted
for.
Both theoretically and observationally,
it has been
shown that increasing irradiation tends to lead to poorer
heat recirculation (e.g. Komacek & Showman 2016;
Schwartz et al. 2017). However, there are a few no-
table exceptions to this rule at high temperatures. Re-
cently, Zhang et al. (2018) reported a heat recirculation
efficiency of ε ∼ 0.2 for the UHJ WASP-33b which is
far higher than would be predicted by theoretical and
observational trends. WASP-12b may also possess an
H2 Dissociation/Recombination in UHJs
7
unusually high heat recirculation efficiency and exhibit
a greater phase offset than would be expected from sim-
ple heat advection1 (Cowan et al. 2012). However, the
power in the second order Fourier series terms from H2
dissociation/recombination seems to make the phase-
curve more sharply peaked and does not seem to be
able to explain the double-peaked phasecurve seen for
WASP-12b by Cowan et al. (2012). Also, while Ar-
cangeli et al. (2018) find evidence of H2 dissociation/
recombination in the atmosphere of WASP-18b, Maxted
et al. (2013) finds the planet has minimal day–night heat
recirculation. Given the expected increase in heat recir-
culation due to H2 dissociation/recombination, this sug-
gests that WASP-18b has only moderate winds and/or
is too cool for these processes to play a strong role in the
heat recirculation of this planet. Finally, near-infrared
observations of KELT-9b, the hottest UHJ currently
known (Gaudi et al. 2017), could provide a fantastic
test of this theory in the very high temperature regime.
T.J.B. acknowledges support from the McGill Space
Institute Graduate Fellowship and from the FRQNT
through the Centre de recherche en astrophysique du
Qu´ebec. The atmospheric model we use in this work is
based upon code originally developed by Diana Jovmir
and Joel Schwartz. We also thank Gabriel Marleau and
Ian Dobbs-Dixon for their helpful insights. We have also
made use of free and open-source software provided by
the Python, SciPy, and Matplotlib communities.
REFERENCES
Amundsen, D. S., Baraffe, I., Tremblin, P., et al. 2014,
de Wit, J., Lewis, N. K., Langton, J., et al. 2016, ApJL,
A&A, 564, A59
820, L33
Angerhausen, D., DeLarme, E., & Morse, J. A. 2015,
PASP, 127, 1113
Dean, J. 1999, Langes handbook of chemistry, 36
Demory, B.-O., de Wit, J., Lewis, N., et al. 2013, ApJL,
Arcangeli, J., Desert, J.-M., Line, M. R., et al. 2018, ArXiv
776, L25
e-prints, arXiv:1801.02489
Bell, T. J., Nikolov, N., Cowan, N. B., et al. 2017, ApJL,
Dobbs-Dixon, I., & Cowan, N. B. 2017, ApJL, 851, L26
Esteves, L. J., De Mooij, E. J. W., & Jayawardhana, R.
847, L2
2013, ApJ, 772, 51
Berardo, D., Cumming, A., & Marleau, G.-D. 2017, ApJ,
Evans, T. M., Sing, D. K., Kataria, T., et al. 2017, Nature,
834, 149
Burrows, A., Sudarsky, D., & Hubeny, I. 2006, ApJ, 650,
1140
Chase, M. W. 1998, J. Phys. Chem. Ref. Data, Monograph
9, 1
Collins, K. A., Kielkopf, J. F., & Stassun, K. G. 2017, AJ,
153, 78
Cowan, N. B., & Agol, E. 2011, ApJ, 729, 54
Cowan, N. B., Fuentes, P. A., & Haggard, H. M. 2013,
MNRAS, 434, 2465
Cowan, N. B., Machalek, P., Croll, B., et al. 2012, ApJ,
747, 82
548, 58
Gaudi, B. S., Stassun, K. G., Collins, K. A., et al. 2017,
Nature, 546, 514
Hebb, L., Collier-Cameron, A., Loeillet, B., et al. 2009,
ApJ, 693, 1920
Heng, K., & Kitzmann, D. 2017, ApJS, 232, 20
Hill, T. L. 1986, An Introduction to Statistical
Thermodynamics (Dover, New York)
Koll, D. D. B., & Komacek, T. D. 2017, ArXiv e-prints,
arXiv:1712.07643
Komacek, T. D., & Showman, A. P. 2016, ApJ, 821, 16
Kreidberg, L., Line, M. R., Bean, J. L., et al. 2015, ApJ,
814, 66
Crossfield, I. J. M., Barman, T., Hansen, B. M. S., Tanaka,
I., & Kodama, T. 2012, ApJ, 760, 140
Lewis, N. K., Knutson, H. A., Showman, A. P., et al. 2013,
ApJ, 766, 95
Dang, L., Cowan, N. B., Schwartz, J. C., et al. 2018, ArXiv
Maxted, P. F. L., Anderson, D. R., Doyle, A. P., et al.
e-prints, arXiv:1801.06548
2013, MNRAS, 428, 2645
1 Depending on the decorrelation method used to reduce the
Spitzer /IRAC data for WASP-12b, the planet either has ε ∼ 0 or
ε ∼ 0.5 (Cowan et al. 2012; Schwartz et al. 2017); although the
former is the preferred model, further observations are critical to
definitively choose between these values and test the predictions
made in this article.
Menou, K. 2012, ApJ, 745, 138
Parmentier, V., Fortney, J. J., Showman, A. P., Morley, C.,
& Marley, M. S. 2016, ApJ, 828, 22
Perez-Becker, D., & Showman, A. P. 2013, ApJ, 776, 134
Pierrehumbert, R. T. 2010, Principles of Planetary Climate
Rauscher, E., & Menou, K. 2010, ApJ, 714, 1334
8
Bell & Cowan
Rink, J. P. 1962, JChPh, 36, 262
Schwartz, J. C., Kashner, Z., Jovmir, D., & Cowan, N. B.
2017, ApJ, 850, 154
Showman, A. P., Fortney, J. J., Lian, Y., et al. 2009, ApJ,
699, 564
Showman, A. P., & Guillot, T. 2002, A&A, 385, 166
Shui, V. H. 1973, JChPh, 58, 4868
Stevenson, K. B., Bean, J. L., Madhusudhan, N., &
Harrington, J. 2014, ApJ, 791, 36
Zhang, M., Knutson, H. A., Kataria, T., et al. 2018, AJ,
155, 83
Zhang, X., & Showman, A. P. 2017, ApJ, 836, 73
|
1806.09160 | 1 | 1806 | 2018-06-24T14:58:22 | Evolution of Titan s high-altitude aerosols under ultraviolet irradiation | [
"astro-ph.EP"
] | The Cassini-Huygens space mission revealed that Titan s thick brownish haze is initiated high in the atmosphere at about 1000 km of altitude, before a slow transportation down to the surface. Close to the surface at altitudes below 130 km, the Huygens probe provided information on the chemical composition of the haze. So far we do not have insights on a possible photochemical evolution of the aerosols composing the haze during their descent. We address here this atmospheric aerosol aging process, simulating in the laboratory how solar vacuum-ultraviolet (VUV) irradiation affects the aerosol optical properties as probed by infrared spectroscopy. An important evolution is found, which could explain the apparent contradiction between the nitrogen-poor infrared spectroscopic signature observed by Cassini below 600 km of altitude in Titan s atmosphere, and a high nitrogen content as measured by the Aerosol Collector and Pyroliser of Huygens probe at the surface of Titan. | astro-ph.EP | astro-ph | Evolution of Titan's high-altitude aerosols under ultraviolet irradiation
Authors: Nathalie Carrasco1,2, Sarah Tigrine1,3, Lisseth Gavilan1, Laurent Nahon3, Murthy S.
Gudipati4
Affiliations
1LATMOS/IPSL, UVSQ, Université Paris-Saclay, UPMC Univ. Paris 06, CNRS, Guyancourt,
France.
2Institut Universitaire de France, France.
3SOLEIL, l'Orme des Merisiers, St Aubin, BP48, F-91192 Gif sur Yvette Cedex, France.
4Science Division, Jet Propulsion Laboratory, Science Division, California Institute of
Technology, 4800 Oak Grove Drive, Pasadena, California 91109, USA.
Correspondence to [email protected]
Abstract:
The Cassini-Huygens space mission revealed that Titan's thick brownish haze is initiated
high in the atmosphere at about 1000 km of altitude, before a slow transportation down to
the surface. Close to the surface at altitudes below 130 km, the Huygens probe provided
information on the chemical composition of the haze. So far we do not have insights on a
possible photochemical evolution of the aerosols composing the haze during their descent.
We address here this atmospheric aerosol aging process, simulating in the laboratory how
solar vacuum-ultraviolet (VUV) irradiation affects the aerosol optical properties as probed
by infrared spectroscopy. An important evolution is found, which could explain the
apparent contradiction between the nitrogen-poor infrared spectroscopic signature
observed by Cassini below 600 km of altitude in Titan's atmosphere, and a high nitrogen
content as measured by the Aerosol Collector and Pyroliser of Huygens probe at the
surface of Titan.
Introduction:
Only one natural satellite in the solar system hosts a dense atmosphere, Titan the largest satellite
of Saturn. With a pressure of about 1.5 bars at the surface, the atmosphere is mainly composed of
molecular nitrogen and methane. The photochemistry of these two molecules produces a
multitude of heavy organic molecules leading to solid aerosols, responsible for the brownish
haze surrounding Titan. The methane concentration is close to 2 % in the stratosphere1 (40-320
km) which contains most of the mass of the organic haze. NASA's Cassini space mission
enabled to observe the aerosols up to 1000 km in Titan's atmosphere thanks to the UVIS
spectrometer 2, revealing that Titan's aerosols are created in the ionosphere, a sub-layer of the
thermosphere where positive and negative ions are produced through solar ultraviolet radiations
and magnetospheric electrons from Saturn 3,4. The large negative ions have been identified as
embryos of aerosols 5. These nanoparticles aggregate and sediment6 leading to the haze observed
by the Cassini Imaging Science Subsystem7 and close to the surface by the Descent Imager and
Spectral Radiometer of the Huygens probe 8.
Pre-Cassini modeling studies suspected a possible aging of the aerosols through UV radiation or
particle impact9,10 without knowing at that time that aerosols embryos were initiated at altitudes
as high as a thousand of kilometers. Yet in the thermosphere highly energetic photons are
present: solar Vacuum-UltraViolet (VUV) flux with wavelengths lower than 200 nm. These
photons could affect the chemical composition of the aerosols since their formation and as a
result the aerosol albedo, impacting the radiative budget of Titan's atmosphere. Our aim is to
address experimentally the key issue of the photochemical aging of the aerosols in the
thermosphere where VUV photons still reach, knowing that the aerosols descent in the
thermosphere, between 1000 and 600 km, lasts one Titanian day (that is about 11 terrestrial days)
6. We will focus on the evolution of their infrared signatures as a sensitive probe of bond
breaking and chemical rearrangement. Those signatures will be compared to the spectra gathered
by the Cassini-VIMS instrument in atmospheric layers below the thermosphere.
2
Results:
The goal of this work is to determine if irradiation at wavelengths representative of the
thermosphere may impact the chemical composition of the aerosols. The photochemical aging
process is simulated in our laboratory experiments by exposing Titan's aerosol analogs to VUV
synchrotron radiation (see Fig. 1).
Several analogues of Titan's aerosols can be considered with different properties and chemical
structures11. In this founding work, we focussed on one type of sample for consistency. The
exploration of the same irradiation process on other possible analogues will be valuable in the
future. We have therefore synthesized a set of samples similar in size and composition to
investigate the effect of the VUV-radiation processing on the freshly formed aerosols. Twenty
film samples with 440 ± 20 nm thicknesses have been prepared during the same experiment on
Si substrates following the protocol detailed in the Methods section.
In the thermosphere aerosols are submitted to solar energetic VUV photons. The absorption by
the atmospheric components leads to an altitude dependent VUV spectrum. Depending on the
altitude where the interaction occurs during their descent, the aerosols will be irradiated with
photons of different major wavelengths. We have chosen two irradiation wavelengths to
experimentally simulate the evolution of the aerosols in the thermosphere: 95 nm representative
of hard photons at about 1000 km and 121.6 nm (Lyman-α) a major VUV contribution in the
solar spectrum penetrating down to 600 km 12. We have obtained these wavelengths at the VUV
DESIRS beamline of SOLEIL synchrotron 13. The total VUV-UV solar flux density at Titan is
about 1014 photons cm-2 s-1 (solar irradiance reference spectra at 1 A.U. 14, attenuated by 0.011
due to the distance from the Sun), whereas the DESIRS beamline provides a photon flux density
of 1016 photons cm-2 s-1. We have therefore chosen irradiation times of a few hours to simulate
the irradiation occurring during one Titan-day.
The evolution of the infrared signatures of the organic films was characterized by ex-situ infrared
absorption spectroscopy in the 1200-3500 cm-1 wavenumber range. The non-irradiated films
have been characterized in earlier publications15. Vibration modes of double bonds C=C and
C=N are found in the 1500-1600 cm-1 range. Triple bonds nitrile -C≡N and isonitrile -N≡C have
their vibration modes peaked at about 2200 cm-1. N-H amine functions show a large absorption
band in the 2700-3700 cm-1 range. And stretching modes of C-H alkyl signatures leads to sharp
bands between 2700 and 3000 cm-1 (-CH3 symmetric and asymmetric and -CH2- asymmetric).
Initial spectra preceding the irradiation experiments were taken for each organic thin film and
normalized at about 1550 cm-1 where the highest absorption value is observed. They are given in
Fig 2 as an envelope capturing the dispersion (twice the standard deviation), and. We focus on
the evolution of three main chemical functions: nitrile bonds R-CN (and isonitrile R-NC) at
~2200 cm-1 (2100-2300 cm-1), C-H bonds at ~2950 cm-1 (2800-3050 cm-1), and amine N-H
bonds at ~3200 cm-1 (3050-3500 cm-1). We find an absorbance dispersion for these three regions
of 5% for nitrile bonds, 15% for C-H bonds, and 10% for N-H bonds.
The effect of VUV irradiation is illustrated in Fig. 2 with the comparison of the absorbance
spectra of a single sample before and after 24 hours irradiation at 121.6 nm. We observe a
general decrease of the targeted signatures, nitrile functions, C-H bonds and N-H bonds. The
time-evolution of this effect is reported in Fig. 3 (upper panel). Given the variability of these
three absorption features, the decrease becomes significant for only 2 hours of irradiation for the
3
nitrile functions, while more than 20 hours of irradiation are needed to induce a modification of
the C-H and N-H bonds.
The samples irradiated at 95 nm evolve similarly, with significant changes observed at shorter
irradiation durations only for the nitrile functional group. The kinetics data of the nitrile decrease
are compared for the two wavelengths irradiations (95 nm and 121.6 nm) in Fig.3. As expected
with the higher energy dose provided at 95 nm, absorption of the nitrile group decreases faster
and more significant with 95 nm irradiation than with 121.6 nm. A loss of 17 ± 5% is obtained at
95 nm after 2 hours of irradiation, which is about twice compared to the one observed at 121.6
nm.
Depending on the wavelength, photons penetrate differently the solid sample. This effect is
quantified by the penetration depth, D(λ). At this layer thickness, the initial radiation intensity is
attenuated by a factor 1/e. The penetration depth of our organic sample is not precisely known.
An approximation can be obtained from the imaginary part k of the VUV refractive index
obtained for another similar material 16. Given the equation
, penetration depths of
D(λ) =
λ
4πk
26 nm and 11 nm are thus expected for radiation at Lyman-α and 95 nm respectively. The
effective layer thickness affected by the radiation is considered at four times the penetration
depth with a residual radiation of 2%. The short penetration depth of VUV photons means that
these photons are effective up to 100 nm of the 440 nm thick sample during VUV irradiation.
Due to the short VUV penetration depth, the material is composed after radiation of two
components: ~20% of the total thickness, which has been irradiated and ~80% which did not
receive VUV photons. Both components contribute to the total infrared transmission signal of the
material after VUV radiation. The experimental 15-20% IR signature attenuation observed on
Figure 3 is similar to the thickness contribution of the irradiated layer, showing that the irradiated
layer possesses totally different IR properties from the non-irradiated layer.
The 95 nm wavelength has been chosen as a typical wavelength below 100 nm simulating the
hard irradiation occurring at about 1000 km of altitude 17. Yet at about 1000 km, the aerosols
initiate their production and reach nanometer-scale dimensions 6,18. The effect observed with
aerosol analogs irradiated at 95 nm will therefore involve a photochemical aging of the aerosols
since the very beginning of their formation process. Below 1000 km, we have shown that the
evolution process is still active by lower energy radiation such as Lyman-α. A strong
modification of Titan aerosol infrared signature is therefore expected through solar VUV aging
all along the aerosols sedimentation in Titan's thermosphere.
4
If we focus on the 2100-2300 cm-1 wavenumber range, an evolution of the shape of the structure
is detected in addition to a change in the intensity. A deconvolution into four gaussian
components enables to characterize this evolution (Fig. 4). Peak spectral position are considered
as free parameters during the gaussian fit procedure in order to take into account possible
changes in the chemical arrangement from the irradiation. They are found centered at 2110,
2140, 2180 and 2240 cm-1 in both cases, with a limited spectral shift between the initial and the
irradiated sample (Table 1). According to previous studies 19,20, these components can be
attributed to aromatic isonitrile (Ar-NC) at 2110 cm-1, aliphatic isonitrile (R-NC) at 2140 cm-1,
conjugated nitrile (C=C-CN) at 2180 cm-1 and aliphatic nitrile (R-CN) at 2240 cm-1. There was a
debate on the component at 2180 cm-1 with a first attribution by Mutsukura and Akita 21 to
isonitrile R-NC structures but, since this wavenumber was found at the upper limit of the
isonitrile region, it is more compatible with a conjugated nitrile signature 19,20. A component for
aromatic nitriles is found at about 2230 cm-1 for similar materials in the literature 19,21, but is
negligible in our case, implying that nitrile termination functions are rarely supported by
aromatic rings.
The integrated areas of the 2110, 2140 and 2240 cm-1 components evolve upon irradiation at
Lyman-α during 24 hrs (see Table 1). The difference observed in the 2100-2300 cm-1 region is
based on a significant decrease of the single component centered at 2180 cm-1, by about 30%.
This component attributed to conjugated nitriles is singularly affected by the VUV irradiation at
Lyman-α and 95 nm. This result is to be compared with the spectral evolution observed after soft
X-ray irradiation of similar aerosol analogs reported in Gavilan et al. 22. The same conjugated
nitrile component at 2180 cm-1 decreased after X-ray exposure, but an aromatization was
simultaneously observed with an important increase of the aromatic nitrile component at 2230
cm-1. This is an important difference between the respective effects of VUV and X-ray
irradiations: aromatization of the conjugated nitriles is much smaller under VUV irradiation. A
decrease of the conjugate nitrile contribution with minimal effect on the other nitrile or isonitrile
signatures in the material involves a loss of this specific nitrile component through a possible
breakage of the C-C bond between the double C=C bond and the nitrile termination function. CN
and/or HCN losses are expected through this process. Note that in the context of Titan's
atmospheric photochemistry, the contribution of X-rays can be neglected, in terms of flux and
absorption cross sections, compared to the photochemistry driven by the VUV photons reaching
the upper atmosphere of Titan.
5
The large 2700-3700 cm-1 signature of the N-H stretching modes of primary and secondary
amine functions decreases after VUV irradiation. This decay is explained by VUV
photodissociation of amines (primary and secondary), initiated by Rydberg excitations. Tertiary
amines are less photolysed as they stabilise by fluorescence. Primary and secondary amines are
therefore depleted, leading to new N-containing functions such as tertiary amines and/or imine.
The production of imine functional groups, C=N, is also consistent with the absorbance increase
in the 1500 cm-1 region. In this process, hydrogen atoms are lost. The consequence is that N-H
amine functions are actually consumed, but with no nitrogen loss. After VUV aging, the samples
remain nitrogen-rich, but amine-poor.
The sharper bands of the C-H signatures in the 2800-3050 cm-1 region show a change in their
distribution after irradiation. To characterise the C-H absorption intensities alone, the underlying
amine absorption band is substracted after interpolation by a linear model in the 2800-3200 cm-1
wavenumber range (Fig. 5). The C-H feature is composed of three main components centered at
2881 cm-1 (CH3 symmetric stretching), 2938 cm-1 (CH2 asymmetric stretching) and 2965 cm-1
(CH3 asymmetric stretching) 19. After 24hrs of irradiation the intensities of the three components
decrease by about 20-30%. This evolution is in agreement with a H-loss of the organic material
through VUV photolysis 23. The decrease is more pronounced for the CH2 groups than the CH3
ones inferring an increase of the CH3/CH2 ratio. This evolution of the CH3/CH2 ratio after
irradiation is relevant to the aerosol signature in Titan atmosphere as measured by the Cassini
Visible and Imaging Mapping spectrometer 24 at altitude ~ 200 km. A slight shift of a few cm-1
also appears for the three components after irradiation typical of a change in the chemical
environment of the C-H bonds. A similar positive shift has been observed in the C-H stretching
signatures of series of molecules, either through steric effects increasing hydrogen-hydrogen
interactions within the material 25 or by the substitution of the adjacent carbon by a nitrogen atom
26. The interactions with both an opposing hydrogen atom or an adjacent nitrogen lone electron
pair hinder the C-H stretch, which requires more energy to occur. The cross-linking of our N-rich
material under VUV irradiation would be consistent with both steric and nitrogen effects. Before
and after irradiation we note the absence of C-H stretching modes supported by aromatic rings
around 3000 cm-1 20. The presence of signatures above 3000 cm-1 in the VIMS data measured at
about 200 km (Fig. 5) is therefore not explained by the VUV processing of the aerosols in
Titan's thermosphere. Other processes should be invoked to explain this feature such as an
aromatic component already present before VUV processing in Titan's thermosphere.
Finally, in the 1500 cm-1 region we find an increase in the absorption of the peak at 1570 cm-1, as
seen in Fig. 2. In this region absorption is due to -C=C- or –C=N- functional groups. We
interpret the increase in this absorption as due to radiation-induced cross-linking of polymers in
the film. This is the only spectral region that showed an increase in the absorption, indicative of
formation of new bonds.
6
Discussion:
The evolution highlighted in this work may also reconcile the contradictory observations on
Titan aerosols made by instruments on Cassini and Huygens at different altitudes in the
atmosphere.
At the surface, the organic haze has been chemically probed by the ACP instrument during the
descent of Huygens (Aerosol Collector and Pyrolyzer). ACP reveals a refractory nucleus rich in
nitrogen 27. The nucleus has been initiated high in the atmosphere: it originates from the
nanometric-scale aerosol embryos produced in the ionosphere. The exact chemical composition
of the embryos is not given by any direct Cassini measurements in Titan ionosphere, but
consistent data support a nitrogen-rich composition involving amine functions. First the fast ion-
driven chemistry occuring in Titan ionosphere builds large nitrogen-rich positive and negative
ions 28-30. Then ammonia is an abundant molecule in Titan ionosphere 31 and a possible important
gas-phase contributor to the formation of amine functions in the aerosols 32. And finally analogs
of Titan ionospheric aerosols produced with plasma discharge in N2-CH4 gas mixtures also
confirm an important amine signatures 15,19.
This aerosol nitrogen content observed at the surface and strongly suspected in the upper
atmosphere seems to disagree with observations in the stratosphere made by infrared
spectroscopy. The aerosols as seen by Cassini-VIMS in the 2900-3000 cm-1 wavenumber region
have mainly a signature of aliphatic C-H bonds 33 and the underlying amine contribution
expected for N-rich aerosols in this wavenumber range remains hardly detectable.
The experimental evidence of the VUV effect on analogs of Titan upper atmospheric aerosols
enables to resolve this apparent contradiction. Chemical growth processes in the upper
atmosphere lead to nitrogen-rich species, as revealed by Cassini INMS 30. The embryos of
aerosols produced in this environment are therefore expected to be N-rich, also in agreement
with our laboratory analogs showing various nitrogen-containing functions: imines, nitriles and
isonitriles, and amines. This work provides evidence that VUV photochemistry with wavelengths
below 150 nm could deplete the sensible primary and secondary amine functions of aerosols in
the upper atmosphere. This would be consistent with the VIMS and CIRS measurements in the
stratosphere, which show no evidence of these functions. The irradiation effect does preserve
nitrogen-bearing functionalities that are more strongly bound in the aerosol skeleton, such as
tertiary amines, imines and nitriles (except conjugated nitriles). This chemical transformation of
the aerosol composition would thus allow the CIRS and VIMS results to agree with the high
nitrogen content reported by the Huygens-ACP instrument close to the surface.
7
References:
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
Niemann, H. B. et al. The abundances of constituents of Titan's atmosphere from the GCMS
instrument on the Huygens probe. Nature 438, 779-784 (2005).
Koskinen, T. T. et al. The mesosphere and lower thermosphere of Titan revealed by
Cassini/UVIS stellar occultations. Icarus 216, 507-534 (2011).
Lavvas, P. et al. Aerosol growth in Titan's ionosphere. Proceedings of the National Academy of
Sciences (2013).
Waite Jr., J. H. et al. The Process of Tholin Formation in Titan's Upper Atmosphere. Science
316, 870-875 (2007).
Coates, A. J. et al. Discovery of heavy negative ions in Titan's ionosphere. Geophys. Res. Lett.
34, L22103 (2007).
Lavvas, P., Sander, M., Kraft, M. & Imanaka, H. Surface Chemistry and Particle Shape:
Processes for the Evolution of Aerosols in Titan's Atmosphere. The Astrophysical Journal 728,
80 (2011).
Porco, C. C. et al. Imaging of Titan from the Cassini spacecraft. Nature 434, 159-168 (2005).
Tomasko, M. G. et al. A model of Titan's aerosols based on measurements made inside the
atmosphere. Planetary and Space Science 56, 669-707 (2008).
Dimitrov, V. & Bar-Nun, A. Aging of Titan's Aerosols. Icarus 156, 530-538 (2002).
Dimitrov, V. & Bar-Nun, A. Hardening of Titan's aerosols by their charging. Icarus 166, 440-443
(2003).
Cable, M. L. et al. Titan Tholins: Simulating Titan Organic Chemistry in the Cassini-Huygens
Era. Chem. Rev. 112, 1882-1909 (2012).
Hörst, S. M. Titan's atmosphere and climate. Journal of Geophysical Research: Planets 122, 432-
482 (2017).
Nahon, L. et al. DESIRS: a state-of-the-art VUV beamline featuring high resolution and variable
polarization for spectroscopy and dichroism at SOLEIL. J. Synchrotron Rad. 19, 508-520 (2012).
Thuillier, G. et al. Solar irradiance reference spectra for two solar active levels. Advances in
Space Research 34, 256-261 (2004).
Gautier, T. et al. Mid- and far-infrared absorption spectroscopy of Titan's aerosols analogues.
Icarus 221, 320-327 (2012).
Khare, B. N. et al. Optical constants of organic tholins produced in a simulated Titanian
atmosphere: From soft x-ray to microwave frequencies. Icarus 60, 127-137 (1984).
Lavvas, P. et al. Energy deposition and primary chemical products in Titan's upper atmosphere.
Icarus 213, 233-251 (2011).
López-Puertas, M. et al. Large Abundances of Polycyclic Aromatic Hydrocarbons in Titan's
Upper Atmosphere. The Astrophysical Journal 770, 132 (2013).
Imanaka, H. et al. Laboratory experiments of Titan tholin formed in cold plasma at various
pressures: implications for nitrogen-containing polycyclic aromatic compounds in Titan haze.
Icarus 168, 344-366 (2004).
Quirico, E. et al. New experimental constraints on the composition and structure of tholins.
Icarus 198, 218-231 (2008).
Mutsukura, N. & Akita, K.-i. Infrared absorption spectroscopy measurements of amorphous CNx
films prepared in CH4/N2 r.f. discharge. Thin Solid Films 349, 115-119 (1999).
Gavilan, L. et al. X-Ray-induced Deuterium Enrichment of N-rich Organics in Protoplanetary
Disks: An Experimental Investigation Using Synchrotron Light. The Astrophysical Journal 840,
35 (2017).
Skurat, V. Vacuum ultraviolet photochemistry of polymers. Nuclear Instruments and Methods in
Physics Research Section B: Beam Interactions with Materials and Atoms 208, 27-34 (2003).
Kim, S. J. et al. Retrieval and tentative indentification of the 3 µm spectral feature in Titan's haze.
Planetary and Space Science 59, 699-704 (2011).
8
Mattioda, A. L. et al. Infrared vibrational and electronic transitions in the dibenzopolyacene
family. Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 130, 639-652
(2014).
Mattioda, A. L. et al. Infrared spectroscopy of matrix-isolated neutral polycyclic aromatic
nitrogen heterocycles: The acridine series. Spectrochimica Acta Part A: Molecular and
Biomolecular Spectroscopy 181, 286-308 (2017).
Israel, G. et al. Complex organic matter in Titan's atmospheric aerosols from in situ pyrolysis and
analysis. Nature 438, 796-799 (2005).
Carrasco, N., Westlake, J., Pernot, P. & Waite Jr, H. in The Early Evolution of the Atmospheres of
Terrestrial Planets 145-154 (Springer New York, 2013).
Vuitton, V. et al. Negative ion chemistry in Titan's upper atmosphere. Planetary and Space
Science 57, 1558-1572 (2009).
Vuitton, V., Yelle, R. V. & McEwan, M. J. Ion chemistry and N-containing molecules in Titan's
upper atmosphere. Icarus 191, 722-742 (2007).
Yelle, R. V. et al. Formation of NH3 and CH2NH in Titan's upper atmosphere. Faraday
Discussions 147, 31-49 (2010).
He, C. & Smith, M. A. A comprehensive NMR structural study of Titan aerosol analogs:
Implications for Titan's atmospheric chemistry. Icarus 243, 31-38 (2014).
Rannou, P. et al. Titan haze distribution and optical properties retrieved from recent observations.
Icarus 208, 850-867 (2010).
9
25
26
27
28
29
30
31
32
33
Acknowledgments:
We are grateful to the general SOLEIL staff for running the facility and providing beamtime
under project number 20120579. J.F. Gil is acknowledged for his technical support and the
development and design of the sample holder. N. Carrasco and L. Gavilan thank the European
Research Council for funding via the ERC PrimChem project (grant agreement No. 636829). We
are grateful to B. Fleury for the preliminary IR spectroscopic measurements, as well as P. Pernot
helpful discussions. We thank M. Béchard for his help and commitment during his short visit in
the team. S. Tigrine acknowledges the University of Paris-Saclay for her thesis funding. MSG's
work at the Jet Propulsion Laboratory, California Institute of Technology was done under a
contract with the National Aeronautics and Space Administration and funded through NASA-
SSW grant "Photochemical Processes in Titan's Atmosphere". L. Nahon acknowledges the
support of the Agence Nationale de la Recherche (ANR-07-BLAN-0293).
Contribution of the authors:
NC supervised the study, participated to the irradiation experiments, treated the infrared
spectroscopic data and drafted the article. All authors discussed the results and commented on
the manuscript. LG characterized the sample thickness by ellipsometric measurements. ST
participated to the irradiation experience. LN conceived the irradiation set-up, prepared the
beamline for the irradiation conditions, and took part in the irradiation experiment. MG was
involved in the analysis of the data, its interpretation, reaction mechanisms, and applications to
Titan's atmosphere.
10
Figures
Fig. 1: Sketch of the experiment where the synchrotron VUV beam irradiates Titan's
aerosol analogs synthesized as thin organic films of 440 ± 20 nm thickness deposited on Si
substrates. Each experiment corresponds to the irradiation of a specific sample at a chosen
wavelength and for a given duration. The 11 mm² surface dimension of the films was
chosen to ensure a flux density of about 1016 ph/sec/cm2 along with a homogenous
irradiation of the sample by the synchrotron beam at the position of the sample holder,
without irradiating the adjacent samples.
11
Fig. 2 : Infrared absorption spectra normalized to the highest absorption value at ~1550
cm-1. The grey envelope shows the dispersion (twice the standard deviation) among the
twenty non-irradiated samples in the 1200-3500 cm-1 range. The red spectrum was
recorded after irradiation for 24h by VUV photons at 121.6 nm on a single sample.
12
Fig. 3: (Above) Irradiation experiments at Lyman-α. Relative decrease with respect to the
original intensity of several infrared band maximum intensities versus irradiation-time.
Hatched boxes indicate the dispersion for non-irradiated samples. Error bars on the data
points correspond to twice the standard deviation. Relative decrease becomes significant
when larger than the dispersion of the non-irradiated samples.
(Below) Decrease of the nitrile intensity at 2182 cm-1 under irradiation at 95 and 121.6 nm
versus irradiation-time. Hatched boxes indicate the dispersion for non-irradiated samples.
Error bars on the data points correspond to twice the standard deviation. Relative decrease
becomes significant when larger than the dispersion of the non-irradiated samples.
13
Fig. 4: Deconvolution into four Gaussian components of the 2050-2300 cm-1 region (above)
before irradiation, (below) after 24 hours of irradiation at Lyman-α. The black line
corresponds to the raw spectrum, the red one to the result from the fitting procedure, the
green lines to the four Gaussian components, and in dotted black line the residual.
14
Fig. 5: Zoom on the aliphatic carbon band after subtraction of the amine contribution:
(Black) before irradiation, (Red) after 24 hrs of irradiation at Lyman-α. (Green) Data
from the Cassini Visible and Imaging Mapping spectrometer (VIMS 24) of Titan's
atmosphere at altitude ~ 200 km.
15
Integrated areas (arbitrary units)
Gaussian component
Shift (cm-1)
Ar-NC @2110 cm-1
R-NC @2140 cm-1
C=C-CN @2180 cm-1
R-CN @2240 cm-1
-1.7
-4,7
-1,2
+3,9
Non-irradiated
0.55±0.03
3.00±0.15
Irradiated
0.65±0.03
3.42±0.17
14.84±0.74
10.84±0.54
2.27±0.11
2.16±0.11
Evolution
+0.1±0.1
+0.4±0.3
-4.0±1.3
-0.1±0.2
Table 1: Effect of the Lyman-alpha irradiation on the integrated areas of the various
vibrational structures of the nitrile bands (see deconvolution in Figure 4). Uncertainties on
integrated areas of the non-irradiated and irradiated samples correspond to the 5%
spectral variability determined in this wavelength range.
16
Methods:
Titan's aerosol analogues
Titan's haze analogs are produced by plasma deposition as thin films onto Silicon windows with
a surface of 11 mm² and a thickness of 0.5 mm. In the thermosphere the methane density profile
varies between 2 and 10% 1. Our synthesis has therefore been done by subjecting a 95-5% N2-
CH4 gas mixture, representative of Titan's average atmospheric composition in the thermosphere
to a radio-frequency electric discharge 2. The organic film production results from the chemistry
induced by electron impact in a N2-CH4 gas mixture using a low pressure of 0.9 mbar
radiofrequency capacitively coupled plasma discharge at room temperature 3,4. The plasma
discharge generates electrons that ionize and dissociate CH4 and N2 similarly to the
magnetospheric charged particles from Saturn 5 and the VUV photons reaching the atmosphere
of Titan 2. The radicals and ions thus produced react allowing for an efficient macromolecular
growth that leads to the production of an organic condensate, which is the laboratory analog of
Titan's atmospheric (or thermosphere) aerosols. These analogs produced with a cold-plasma
technique are the one simulating best the limited results obtained by the ACP instrument on
ESA-Huygens when analyzing the refractory nucleus of the aerosols close to Titan surface 6,7.
This result suggests the importance of reproducing in the laboratory analogs the nitrogen content
detected in Titan's aerosols. Previous analysis by UV/VUV spectroscopies 8,9 and solid NMR 10
characterized the chemical structure of our analogs. Those are mainly based on C-N unsaturated
bonded units, of which triazine (C3N3) is an important present component. The samples are
considered as analogs of Titan's atmospheric aerosols in terms of replication of the energy
source and gas phase composition, but were produced with a higher pressure and temperature
than on Titan. The general relevance of laboratory simulations is described in detail in a review
paper by Cable et al. 5.
The samples are slightly contaminated by oxygen (~2% of the total elementary composition 11),
in spite of the strict cleaning protocol described in 4 and based on baking under vacuum followed
by a pure N2 plasma cleaning. This contamination occurs on the surface of the sample when
exposed to ambient air 12. However this contamination is not an issue here, as the samples
infrared spectra treated here are not modified by the small oxygen content, considering the
intense signatures of the C-C and C-N bands in the samples. Another source of contamination
could be considered during the irradiation process in the vacuum chamber. Radicals and ions
released from a sample during its irradiation could interact at the surface of the non-irradiated
samples adjacent on the sample holder. No significant thickness evolution is observed after
irradiation, showing a negligible quantitative loss or addition of solid material after irradiation. If
occurring, a superficial nanometer-scale contamination would not be detected by the infrared
transmission spectroscopy diagnosis and is neglected in the present study.
A film thickness of 440 ± 20 nm is found by spectroscopic ellipsometry in the 370–1000 nm
spectral range (M-2000V ellipsometer, J.A. Woollam Co) 13. We verify that the films are similar
in size and composition prior the irradiation experiment by comparing their IR signatures. We
also make sure that they are homogeneous by recording the IR signatures on different 400 µm
×400 µm spots of the same sample.
17
VUV irradiation of the samples
Considering a solar irradiance reference spectra at 1 A.U. 14, attenuated by 0.011 due to the
distance from the Sun, the total VUV-UV solar flux density reaching Titan is about 1014 photons
cm-2 s-1, and the photon flux density at Lyman-α is 3×109 photons cm-2 s-1.
The irradiation duration of the aerosols in the thermosphere corresponds to the fraction of time
spent by the aerosols between 1000 and 600 km during which the aerosols are irradiated by the
sun. Microphysical models predict that the residence time of the aerosols in Titan thermosphere
is about 11 days (106 s), between their formation at 1000 km and their aggregation in fractal
dimensions triggered at 600 km 15. As a Titan-day lasts about 16 days (1.4 x 106 s), the aerosols
are irradiated most of the time during their sedimentation in Titan thermosphere. Considering
Titan's rotation and the variation of insolation over Titan's disk, the irradiation time over one
day on Titan has to be divided by about a factor of four. During their descent in Titan
thermosphere corresponding to about one Titan-day, aerosols receive therefore a total VUV
irradiation dose of about 3.5×1019 photons cm-2.
In our experiments, the 11 mm² sample surface is entirely illuminated by the beam spot of the
quasi-monochromatic DESIRS-synchrotron beamline 16. The samples are positioned vertically
on a sample holder at 107 cm from the DESIRS beamline focal point. A high precision
manipulator enables to translate from one sample to another. Each sample on the sample holder
is irradiated with a specific wavelength (95 or 121.6 nm) and a chosen irradiation-duration. The
samples are submitted to the irradiation of the DESIRS monochromator at SOLEIL synchrotron.
A gas filter, filled with xenon or argon according to the irradiation wavelength, eliminates the
high harmonics of the undulator 17. In order to maximize the irradiation dose, the grating mirror
is settled at the 0th order position, where it behaves as a mirror. At a given energy E, the pseudo-
Gaussian spectral band of the undulator is transmitted with a spectral resolution ΔE/E = 7% at
half maximum height. With this configuration, the photon flux density is two orders of
magnitude larger than the corresponding UV solar flux density. The photon flux density remains
low enough to ensure single-photon interactions with the material. The dose is adjusted by
choosing shorter irradiation times of typically a few hours in our experiments. For example
during a 24h synchrotron experiment, samples are exposed to a total irradiation dose of 8.6×1020
photons cm-2, comparable to the total VUV irradiation dose received by the thermospheric
aerosols during one Titan-day. The total allocated synchrotron time (4 days) has been distributed
in order to measure time-dependent evolutions of the samples under irradiation at both
wavelengths. Three samples were irradiated at Lyman-α during 3, 10 and 24h respectively,
whereas four samples were irradiated at 95 nm during 30 min, 1, 3 and 24h respectively.
Infrared absorption spectroscopy
The evolution of the infrared signatures of the samples was characterized by ex-situ mid-IR
absorption spectroscopy in the 1200-3500 cm-1 wavenumber range. Initial spectra preceding the
irradiation experiments were taken for each organic thin film. All pre- and post-irradiation
infrared measurements were performed with the Thermo Scientific Nicolet iN10 MX
spectrometer at the SMIS beamline of the synchrotron SOLEIL facility. We used a mercury
cadmium telluride (MCT) detector for a transmission analysis and spectral resolution of 4 cm-1.
18
Data availability
The data that support the plots within the paper and other findings are available from the
corresponding author upon reasonable request.
Additional references
1
Waite Jr, J. H. et al. Ion Neutral Mass Spectrometer Results from the First Flyby of Titan. Science
308, 982-986 (2005).
Szopa, C., Cernogora, G., Boufendi, L., Correia, J.-J. & Coll, P. PAMPRE: A dusty plasma
experiment for Titan's tholins production and study. Planet. Space Sci. 54, 394-404 (2006).
Gautier, T. et al. Mid- and far-infrared absorption spectroscopy of Titan's aerosols analogues.
Icarus 221, 320-327 (2012).
Sciamma-O'Brien, E., Carrasco, N., Szopa, C., Buch, A. & Cernogora, G. Titan's atmosphere: An
optimal gas mixture for aerosol production? Icarus 209, 704-714 (2010).
Cable, M. L. et al. Titan Tholins: Simulating Titan Organic Chemistry in the Cassini-Huygens
Era. Chem. Rev. 112, 1882-1909 (2012).
Coll, P. et al. Can laboratory tholins mimic the chemistry producing Titan's aerosols? A review in
light of ACP experimental results. Planetary and Space Science 77, 91–103 (2013).
Israel, G. et al. Complex organic matter in Titan's atmospheric aerosols from in situ pyrolysis and
analysis. Nature 438, 796-799 (2005).
Quirico, E. et al. New experimental constraints on the composition and structure of tholins.
Icarus 198, 218-231 (2008).
Cunha de Miranda, B. et al. Molecular Isomer Identification of Titan's Tholins Organic Aerosols
by Photoelectron/Photoion Coincidence Spectroscopy Coupled to VUV Synchrotron Radiation.
The Journal of Physical Chemistry A 120, 6529-6540 (2016).
Derenne, S. et al. New insights into the structure and chemistry of Titan's tholins via 13C and 15N
solid state nuclear magnetic resonance spectroscopy. Icarus 221 844–853 (2012).
Fleury, B. et al. Influence of CO on Titan atmospheric reactivity. Icarus 238, 221-229 (2014).
Carrasco, N., Jomard, F., Vigneron, J., Etcheberry, A. & Cernogora, G. Laboratory analogues
simulating Titan's atmospheric aerosols: Compared chemical compositions of grains and thin
films. Planetary and Space Science 128, 52-57 (2016).
Mahjoub, A. et al. Influence of methane concentration on the optical indices of Titan's aerosols
analogues. Icarus 221, 670-677 (2012).
Thuillier, G. et al. Solar irradiance reference spectra for two solar active levels. Advances in
Space Research 34, 256-261 (2004).
Lavvas, P., Sander, M., Kraft, M. & Imanaka, H. Surface Chemistry and Particle Shape:
Processes for the Evolution of Aerosols in Titan's Atmosphere. The Astrophysical Journal 728,
80 (2011).
Nahon, L. et al. DESIRS: a state-of-the-art VUV beamline featuring high resolution and variable
polarization for spectroscopy and dichroism at SOLEIL. J. Synchrotron Rad. 19, 508-520 (2012).
Mercier, B. et al. Experimental and theoretical study of a differentially pumped absorption gas
cell used as a low energy-pass filter in the vacuum ultraviolet photon energy range. J. Vac. Sci.
Tech. A 18, 2533-2541 (2000).
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
19
|
1803.06069 | 1 | 1803 | 2018-03-16T03:41:30 | A Preliminary Analysis of the Shangri-La Bolide on 2017 Oct 4 | [
"astro-ph.EP"
] | At 12:07 UT (8:07 pm China Standard Time) on 2017 Oct 4, a bright bolide was widely observed in the Shangri-La region in the Province of Yunnan, China (Figure 1). The event was well observed by the general public as it took place on the night of the Mid Autumn Festival which associates with moon gazing. Sonic booms and ground shaking were reported in an area about a thousand square kilometers wide to the northwest of the Shangri-La City. Data from the U.S. government sensor suggested that the impact energy of the event is approximately 0.54 kt TNT equivalent, with the terminus of the bolide positioned at $28.1^\circ$ N, $99.4^\circ$ E. This is the largest observed bolide event overland since the bolide event took place in Mauritania on 2016 Jun 27 (1.2 kt). In this Research Note we present a preliminary analysis of this event. | astro-ph.EP | astro-ph | Draft version March 19, 2018
Typeset using LATEX RNAAS style in AASTeX62
8
1
0
2
r
a
M
6
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
6
0
6
0
.
3
0
8
1
:
v
i
X
r
a
A Preliminary Analysis of the Shangri-La Bolide on 2017 Oct 4
Quan-Zhi Ye (叶泉志)1, 2
1Division of Physics, Mathematics and Astronomy, California Institute of Technology, Pasadena, CA 91125, U.S.A.
2Infrared Processing and Analysis Center, California Institute of Technology, Pasadena, CA 91125, U.S.A.
(Received –; Revised –; Accepted –)
Keywords: meteorites, meteors, meteoroids
At 12:07 UT (8:07 pm China Standard Time) on 2017 Oct 4, a bright bolide was widely observed in the Shangri-La
region in the Province of Yunnan, China (Figure 1). The event was well observed by the general public as it took place
on the night of the Mid Autumn Festival which associates with moon gazing. Sonic booms and ground shaking were
reported in an area about a thousand square kilometers wide to the northwest of the Shangri-La City. Data from the
U.S. government sensor suggested that the impact energy of the event is approximately 0.54 kt TNT equivalent, with
the terminus of the bolide positioned at 28.1◦ N, 99.4◦ E. This is the largest observed bolide event overland since the
bolide event took place in Mauritania on 2016 Jun 27 (1.2 kt).
International Meteor Organization and American Meteor Society operates fireball report programs that col-
lect world-wide observations of fireballs and bolides (https://www.imo.net/observations/fireballs/fireballs/ and
https://www.amsmeteors.org/fireballs/, see Hankey et al. 2014; Hankey & Perlerin 2015). However, very few (2)
reports of the Shangri-La event have been archived in either database. Here we collect various of accounts from Weibo
users, forum posts, and media reports, tabulated in Table 1.
Table 1. Compilation of the visual and video accounts of the Shangri-La bolide. Some accounts are extracted from the post
by Liu Wenjieb.
Site
Coordinate
Concurrent
Delayed
Shaking?
Source
Note
boom?
boom?
Balagezong National Park
(巴拉格宗国家公园)
Baoshan City (保山)
Benzilan (奔子栏)
Dali City (大理)
..
Hongpo Village (红坡村)
Lijiang City (丽江)
Niding Village (尼丁村)
Nixi Township (尼西县城)
Nujiang City (怒江)
Ramwok, Tibet (然乌)
Shusong Village (书松村)
Tangman Village (汤满村)
Xiaruo Township (霞若乡)
Xingfu Village (幸福村)
Shangri-La City
(Zhongdian/中甸)
28.285◦ N, 99.432◦ E
25.1◦ N, 99.2◦ E
28.24◦ N, 99.30◦ E
25.6◦ N, 100.3◦ E
25.535◦ N, 100.330◦ E
27.81◦ N, 99.81◦ E
25.964◦ N, 100.157◦ E
28.19◦ N, 99.24◦ E
28.1◦ N, 99.5◦ E
25.8◦ N, 98.9◦ E
29.317◦ N, 96.990◦ E
28.27◦ N, 99.19◦ E
28.02◦ N, 99.49◦ E
27.80◦ N, 99.30◦ E
28.14◦ N, 99.43◦ E
27.8◦ N, 99.7◦ E
X
X
X
X
X
X
X
X
X
X
X
X
Media report
"Passing overhead and heading E, red"
X
X
Witness report
"Sonic boom and shaking like a landslide is taking place"
Media report
Multiple accounts
X
X
X
X
X
X
X
X
X
X
X
X
Media report
Multiple accounts
Dashcam
G56 freeway facing NW
Media report
Dashcam
Liu's post
G5611 freeway facing NW
House shaking, like a gas explosion
Media report
"Pigs run out of barn"
Media report
Multiple accounts
S201 highway facing E
Dashcam
Liu's post
Media report
Liu's post
Liu's post
Witness report
ahttp://bbs.tianya.cn/post-travel-821029-1.shtml, in Chinese. Retrieved 2017 Dec 19.
bhttp://bbs.tianya.cn/post-travel-821029-1.shtml, in Chinese. Retrieved 2017 Dec 19.
Figure 2 shows that moderate ground shaking was limited to an area about 30 × 30 km2 around Hongpo Village,
∼ 40 km northwest of the Shangri-La City (labeled as Deqen/迪庆藏族自治州in the figure). Curiously, reports from
Corresponding author: Quan-Zhi Ye
[email protected]
2
Ye
Figure 1. Location of the Shangri-La bolide.
Hongpo Village did not indicate ground shaking. It is not clear whether this was the case or the information was
simply omitted. According to the local government agency, no significant property damage or loss has so far been
reported1.
We also collected 3 dashboard camera footages that the shooting locations are precisely known: two from freeways
near the cities of Lijiang and Dali in Yunnan, approximately 100 and 300 km to the south, one from Rawu, Tibet,
∼300 km to the northwest (Figure 3). Initial analysis shows that the trajectory and motion of the bolide broadly agree
the U.S. government sensor data. The Rawu footage included the full moon in the field of view, which appeared to be
much less bright than the bolide at its peak brightness (Figure 3, frame b).
The event was also detected seismically and infrasonically. China Earthquake Agency has reported positive detection
by two seismic stations, Zhongdian (27.824◦ N, 99.707◦ E) and Gongshan (27.740◦ N, 98.665◦ E), registered at a Richter
scale of 2.12. The CTBTO Kunming station and an infrasound array deployed in western Yunnan also detected the
event (W. Su, private communication). Both datasets are not yet publicly available.
After the termination of ablation, meteorites decelerate to a speed of a few km/s and will typically spend a few
minutes airborne before reaching the ground (Ceplecha et al. 1998). Atmospheric sounding data obtained by the
nearby Xichang Station in Sichuan, ∼250 km to the east3, revealed that the local atmosphere was dominated by
westerly wind, largely parallel to the travel direction of the bolide. Therefore, we expect that any surviving meteorites
would show little transverse motion. We estimate that the possible strewn field to be ∼ 100 km northeast of the bolide
terminus near the Yunnan–Sichuan border (Figure 4). This is in line with independent calculation carried out by R.
Matson (R. Matson, private communication).
1 http://news.chinaxiaokang.com/dujiazhuangao/2017/1006/260312.html, in Chinese. Retrieved 2017 Dec 19.
2 http://news.xinhuanet.com/tech/2017-10/06/c 1121765328.htm, in Chinese. Retrieved 2017 Dec 19.
3 Available
from an archive maintained by the Department of Atmospheric Science at
the University of Wyoming,
http://weather.uwyo.edu/upperair/sounding.html, retrieved 2018 Jan. 4.
A Preliminary Analysis of the Shangri-La Bolide on 2017 Oct 4
3
Figure 2. Bolide witness and ground shaking reports in the Shangri-La area. Shangri-La City is labeled as Deqen/迪庆藏族自
治州in the figure. Also marked is the fireball detection reported by the U.S. Department of Defense (DoD) sensor.
We also attempted to use the Doppler weather radar data to look for micrometeorites deposits in the atmosphere.
However, the predicted region is not covered by the Chinese weather radar network; the closest Doppler radar at
Lijiang is about 200 km to the south with an effective observing range of 150 km.
It was reported that "several hundreds of" meteorite hunters had arrived on the scene to search for meteorites4.
Many chose to search in the northwest of Shangri-La City where most of the shock wave reports surfaced. Both this
region and our estimated strewn field is very mountainous, but the latter is more so and is sparsely populated. At the
time of this writing, the searches are ceased5 and there is no credible successful recovery of the Shangri-La meteorite.
Unlike the U.S. and Europe, China is yet to establish a dedicated government-level body that oversees the research
of near-Earth objects. Bolides that are energetic enough to be recorded by the U.S. government sensor occur in the
country about once per year, with the last ones being the 2014 Xilin Gol event and the 2015 Gansu event. The Xilin
Gol event, took place in the early morning of 2014 Nov 5 over the Gobi desert ∼ 300 km north of Beijing, was of
similar magnitude to the Shangri-La event and had also caught considerable public attention. It was believed that
multi-kilogram meteorites from the Xilin Gol event had reached the ground. Efforts (mostly led by amateurs) were
launched to recover these meteorites but no successful recovery has been reported so far.
In a global scale, meteorite-dropping events are quite frequent. Using the event rate derived by Halliday et al. (1984)
we estimate that event with > 1 kg meteorite fall occurs once per week on average over the entire Chinese territory.
Amateur astronomers in several parts of China have recently begun to build video camera networks, in the hope to
better constrain the trajectory of the next meteorite-dropping events. As of early 2018, camera networks in Beijing,
Shandong, Guangdong, northern Xinjiang and northwestern Tibet are operational (Ye 2018).
4 http://news.sina.com.cn/o/2017-10-10/doc-ifymrcmm9945166.shtml, in Chinese. Retrieved 2018 Jan 3.
5 http://news.sina.com.cn/s/wh/2017-10-22/doc-ifymyyxw4058501.shtml, in Chinese. Retrieved 2018 Jan 3.
4
Ye
Figure 3. Footage of the dashboard camera provided by Tan Kaixin, who was then traveling on an eastbound vehicle on
Highway S201 near Rawu, Tibet. The clock of the camera was not calibrated and the timings on the lower-right corner are
likely inaccurate.
I thank Peter Brown and Rob Matson for helpful discussions, as well as all the witnesses for sharing their videos
and accounts on the internet. The author is supported by the GROWTH project (National Science Foundation Grant
No. 1545949).
A Preliminary Analysis of the Shangri-La Bolide on 2017 Oct 4
5
Figure 4. Estimated strewn field of the Shangri-La meteorites.
REFERENCES
Ceplecha, Z., Borovicka, J., Elford, W. G., et al. 1998,
Hankey, M., & Perlerin, V. 2015, in International Meteor
SSRv, 84, 327, doi: 10.1023/A:1005069928850
Halliday, I., Blackwell, A. T., & Griffin, A. A. 1984,
Science, 223, 1405, doi: 10.1126/science.223.4643.1405
Conference Mistelbach, Austria, ed. J.-L. Rault &
P. Roggemans, 192–196
Hankey, M., Perlerin, V., Lunsford, R., & Meisel, D. 2014,
in Proceedings of the International Meteor Conference,
Poznan, Poland, 22-25 August 2013, ed. M. Gyssens,
P. Roggemans, & P. Zoladek, 115–119
Ye, Q.-Z. 2018, Astronomical Research & Technology.
https://arxiv.org/abs/1708.00139
|
1702.08618 | 1 | 1702 | 2017-02-28T03:05:41 | A Volcanic Hydrogen Habitable Zone | [
"astro-ph.EP"
] | The classical habitable zone is the circular region around a star in which liquid water could exist on the surface of a rocky planet. The outer edge of the traditional N2-CO2-H2O habitable zone (HZ) extends out to nearly 1.7 AU in our Solar System, beyond which condensation and scattering by CO2 outstrips its greenhouse capacity. Here, we show that volcanic outgassing of atmospheric H2 on a planet near the outer edge can extend the habitable zone out to ~2.4 AU in our solar system. This wider volcanic hydrogen habitable zone (N2-CO2-H2O-H2) can be sustained as long as volcanic H2 output offsets its escape from the top of the atmosphere. We use a single-column radiative-convective climate model to compute the HZ limits of this volcanic hydrogen habitable zone for hydrogen concentrations between 1% and 50%, assuming diffusion-limited atmospheric escape. At a hydrogen concentration of 50%, the effective stellar flux required to support the outer edge decreases by ~35% to 60% for M to A stars. The corresponding orbital distances increase by ~30% to 60%. The inner edge of this HZ only moves out by ~0.1 to 4% relative to the classical HZ because H2 warming is reduced in dense H2O atmospheres. The atmospheric scale heights of such volcanic H2 atmospheres near the outer edge of the HZ also increase, facilitating remote detection of atmospheric signatures. | astro-ph.EP | astro-ph | Accepted in The Astrophysical Journal Letters
A VOLCANIC HYDROGEN HABITABLE ZONE
Ramses M. Ramirez1,2, and Lisa Kaltenegger1,3
1Carl Sagan Institute, Cornell University, Ithaca, NY
2Cornell Center of Astrophysics and Planetary Science, Cornell University, Ithaca, NY
3Department of Astronomy, Cornell University, Ithaca, NY
Corresponding author - Ramses Ramirez email: [email protected], phone 480-296-6477
ABSTRACT
The classical habitable zone is the circular region around a star in which liquid water could
exist on the surface of a rocky planet. The outer edge of the traditional N2-CO2-H2O habitable zone
(HZ) extends out to nearly ~1.7 AU in our Solar System, beyond which condensation and
scattering by CO2 outstrips its greenhouse capacity. Here, we show that volcanic outgassing of
atmospheric H2 on a planet near the outer edge can extend the habitable zone out to ~2.4 AU in
our solar system. This wider volcanic hydrogen habitable zone (N2-CO2-H2O-H2) can be sustained
as long as volcanic H2 output offsets its escape from the top of the atmosphere. We use a single-
column radiative-convective climate model to compute the HZ limits of this volcanic hydrogen
habitable zone for hydrogen concentrations between 1% and 50%, assuming diffusion-limited
atmospheric escape. At a hydrogen concentration of 50%, the effective stellar flux required to
support the outer edge decreases by ~35% to 60% for M - A stars. The corresponding orbital
distances increase by ~30% to 60%. The inner edge of this HZ only moves out ~0.1 to 4% relative
to the classical HZ because H2 warming is reduced in dense H2O atmospheres. The atmospheric
scale heights of such volcanic H2 atmospheres near the outer edge of the HZ also increase,
facilitating remote detection of atmospheric signatures.
Key words: planets - habitable zones - atmospheres - hydrogen - carbon dioxide -volcanism
1
1. INTRODUCTION
The habitable zone (HZ) is the circular
region around a star(s) in which liquid water
could be stable on a rocky planet's surface
(e.g. Kasting et al., 1993, Kaltenegger &
Haghighipour
2013, Haghighipour &
Kaltenegger 2013) and possible atmospheric
biosignatures may be detected
(e.g.
Kaltenegger et al. 2010). The classical N2-
CO2-H2O HZ is defined by the greenhouse
effect of CO2 and H2O vapor. The outer edge,
the CO2 maximum greenhouse limit, is
defined as
the distance beyond which
condensation and scattering by CO2 outstrips
its greenhouse capacity. The inner edge is
defined where mean surface temperatures
exceed the critical point of water (~647 K,
220 bar), triggering a runaway greenhouse
that leads to rapid water loss to space on very
short timescales (see Kasting et al. 1993 for
details). Additional greenhouse gases could
further extend the HZ. It had been suggested
that young orphan planets unbound to their
host stars could accrete massive primordial
hydrogen envelopes
(Stevenson 1999).
Models show that a super-Earth with a 40 bar
primordial hydrogen atmosphere could
maintain
surface
temperatures out to 10 AU from a solar-type
G-star (Pierrehumbert and Gaidos 2011). The
potency of this greenhouse effect arises from
collision-induced absorption (CIA) caused
by self-broadening from H2-H2 collisions,
allowing H2 to function noncondensibly out
to great distances (Pierrehumbert and Gaidos,
2011). For comparison, condensation and
scattering effects at high CO2 partial
pressures limit the outer edge of the classical
HZ in our solar system to ~1.7 AU (e.g.
Kasting et al., 1993).
above-freezing mean
However, maintaining large amounts
of primordial hydrogen over geological
timescales is difficult (Pierrehumbert and
Gaidos, 2011). Hydrogen is a very light
greenhouse gas, and without a continuous
source, hydrodynamic escape to space would
strip a 5 Earth-mass habitable zone rocky
planet with 50 bars of primordial atmospheric
H2 within just a few million years. Primordial
hydrogen lifetimes in atmospheric models
only extend to ~100 million years at H2
partial pressures of several hundred bar
(Wordsworth, 2012).
partial
pressures
Another way to generate atmospheric
hydrogen on terrestrial planets is through
volcanism. Paleoclimate studies of early
Earth and Mars show that volcanism could
have outpaced escape of H2 on both planets,
maintaining clement conditions with modest
H2
bar
(Wordsworth and Pierrehumbert, 2013;
Ramirez et al., 2014). Reduced mantle
conditions could have favored enhanced
outgassing of H2 over longer timescales. In
these models hydrogen is not the major
atmospheric constituent and is continuously
replenished by volcanism that offsets losses
to space.
these volcanic hydrogen
atmospheres, it is foreign-broadening by the
remaining background atmosphere
that
excites roto-translational bands within the
hydrogen, increasing greenhouse warming in
spectral regions where CO2 and H2O absorb
poorly (Ramirez et al. 2014).
below
1
In
In this work, we assess the impact that
foreign broadening of volcanic hydrogen in a
planet's atmosphere has on the classical N2-
CO2-H2O habitable zone (Kasting et al.,
1993; Kopparapu et al., 2013). We derive the
limits of this volcanic-hydrogen HZ (N2-
CO2-H2O-H2) for host stars with Teff ranging
from 2,600K to 10,000 K.
2
2. METHODOLOGY
2.1 Climate modeling procedures
2014;
Ramirez
10,000K
(Ramirez
As in previous studies of the HZ (e.g.
Kopparapu et al., 2013; Ramirez and
Kaltenegger,
and
Kaltenegger, 2016) we used a 1-D radiative-
convective climate model
to compute
habitable zone boundaries for stars of stellar
effective temperatures (Teff) ranging from
2,600K
and
Kaltenegger, 2016). The model uses
correlated-k coefficients
to parameterize
absorption by H2O and CO2 across 38 solar
intervals and 55 infrared intervals (e.g.
Kopparapu et al. 2013 for details). We model
CO2-H2 collision-induced absorption (CIA)
using
initio
calculations (Wordsworth et al., 2017), CO2-
CO2 CIA using established parameterizations
(Gruszka and Borysow, 1997; Gruszka and
Borysow, 1998; Baranov et al., 2004), self-
broadening of H2-H2 pairs (Borysow, 2002)
and N2 foreign-broadening (see Ramirez et
al. 2014 for details).
computed
to
recently
ab
it
flux
required
the mean
calculations,
to maintain
We use inverse calculations, where
the surface temperature is specified and the
solar
is
calculated, to determine the limits of the HZ
(following e.g., Kasting et al., 1993;
Kopparapu et al., 2013). For the outer edge
HZ
surface
temperature is fixed at 273 K, above which
liquid water is stable on a planetary surface.
Stratospheric temperatures are fixed at 154
K, which is representative for a planet near
the outer edge of the habitable zone (e.g.
Kopparapu et al. 2013). For the inner edge,
the stratospheric temperature is 200 K and the
surface temperature is gradually increased
from a starting temperature of 200 K, which
simulates pushing the planet closer and closer
to the star until a runaway greenhouse is
triggered (e.g. Kasting et al., 1993).
3
conditions
The input of hydrogen from volcanic
sources in our models is balanced by its
escape to space. Actual hydrogen escape
rates will be planet-specific, therefore we
assume that hydrogen escapes at the diffusion
limit, which is the fastest that H can escape at
the concentrations considered (e.g Walker,
1977). Thus, our results may overestimate
hydrogen escape and thus provide a lower
bound on H concentrations. As with previous
HZ studies, we assume planets support plate
tectonics, including a carbonate-silicate cycle
capable of maintaining habitability over long
timescales (e.g. Kasting et al., 1993, Ramirez
& Kaltenegger 2016). We also assume
on
highly-reduced mantle
terrestrial planets, as
from
meteoritic evidence from early Mars (e.g.
Grott et al., 2011). Hydrogen sources and
sinks need to balance to support a continuous
hydrogen concentration. Our model planet
has the same outgassing rates per unit area as
those on our present day planet based on the
inference that the heat flux of early Mars
could have been similar to that of modern
Earth (Montési and Zuber, 2003; Ramirez et
al., 2014). This allows us to constrain typical
H2 concentrations for reducing mantles of
outer edge planets to ~ 1 – 30% (Ramirez et
al., 2014; Batalha et al., 2015). For
comparison we
extreme
concentration of 50% (assuming even higher
outgassing rates or less efficient hydrogen
escape) intended to simulate potentially
higher outgassing rates (e.g. on hypothetical
planets with unusually high geothermal
fluxes).
include one
suggested
Each of our model atmospheres
contains 1 bar of N2 and H2 concentrations of
1, 5, 10, 20, 30, and 50%, respectively, of the
dry (N2+CO2+H2) atmosphere. Oxygen has
only a weak greenhouse effect and is replaced
by N2
these calculations (following
Kasting et al. 1993). For the outer edge
calculations we vary the CO2 partial pressure
in
from 1x10-2 to 34.7 bar, the saturation CO2
partial pressure at 273K. CH4 is not expected
to build up to appreciable concentrations in
CO2-H2 atmospheres because of its short
photochemical lifetime when H2 is not the
major constituent (Bains et al. 2014). Recent
work shows that CH4 would likely oxidize to
CO2 in warm CO2-H2 atmospheres (Batalha
et al., 2015).
At the inner edge, our volcanic-H2
atmospheres contain 1 bar of N2 and 330 ppm
CO2 (following Kasting et al., 1993). As in
the case of the outer edge, we model H2
concentrations that are 1, 5, 10, 20, 30 and
50% of the dry (N2+CO2+H2) atmosphere.
We gradually increase planetary surface
temperatures from 200K to the critical point
of water (647.1K, 220.6 bars H2O), which
leads to the runaway greenhouse and rapid
evaporation of the surface water inventory
(e.g., Kasting et al., 1993). The high
atmospheric H2O mole fraction at the inner
edge reduces the warming effect of H2. Note
that the H2 amount is scaled to the dry
atmosphere because volcanism
is not
expected to increase as H2O vapor mixing
ratios increase.
In keeping with previous work (see
e.g. Kopparapu et al. 2013) all model
atmospheres are fully-saturated: A moist CO2
adiabat is followed in the upper atmosphere
for the outer edge calculations. A moist H2O
adiabat is followed in the upper atmosphere
for the inner edge model calculations. We
also employ six solar zenith angles and a
surface albedo of 0.31 to mimic the combined
clouds and
reflection, which
reproduces the mean surface temperature for
present Earth (ibid). Because there is no self-
consistent cloud model for 1D HZ studies
(see Zsoms et al. 2012), surface and cloud
albedo are held constant for all simulations
(following previous 1-D HZ studies, see e.g.
Kasting et al., 1993; Kopparapu et al., 2013),
even though cloud feedback is expected to
surface
4
for
atmospheric
compositions
change
different from those of present-day Earth.
However, our understanding of clouds is still
evolving and their applicability to water-rich
atmospheres (i.e. inner edge of the HZ) is not
fully understood yet. Note that overall
planetary albedo changes as the atmospheric
composition is varied.
3. RESULTS
Adding volcanic hydrogen moves both
the inner and the outer edges of the classical
HZ outward and widens it. The incident
stellar fluxes (Seff) corresponding to the outer
edges of the classical and new volcanic-H2
HZ, are shown in fig.1. For comparison, an
alternative HZ limit that is not based on
atmospheric models (i.e. classical HZ), but
on empirical observations of our solar
system, is also shown in figs. 1-2. The inner
edge of this empirical HZ is defined by the
stellar flux received by Venus when we can
exclude the possibility that it had standing
water on the surface (about 1Gyr ago),
equivalent to a stellar flux of Seff=1.77
(Kasting et al.1993). The outer edge is
defined by the stellar flux that Mars received
at the time that it may have had stable water
on its surface (about 3.8 Gyr ago), which
corresponds to Seff=0.32 (ibid). The incident
flux in both figures is normalized to that
received by Earth's orbit (~1360 W/m2), S0.
The incident stellar flux that is needed to
maintain liquid water surface temperatures at
the outer edge of our solar system decreases
from 0.3576 to 0.267 (25% decrease) and to
0.201 (44% decrease) with 5% and 30% H2,
respectively (fig.1). This corresponds to
increases in the orbital distance of the
maximum greenhouse limit of CO2 at the
outer edge HZ from 1.67 AU to 1.94 AU and
2.23 AU, respectively (16% and 34%
increases, respectively; fig.2). With 50% H2
the limits extend farther, corresponding to an
Seff of 0.172 (52% increase) and an outer edge
distance of 2.4 AU (44% increase) in our own
Solar System. Overall, the Seff required to
support the outer edge decreases by a
maximum of ~35 – 60% (at 50% H2) for M -
A star types, resulting in respective distance
increases of ~30 to 60% (fig. 2).
Figure 1: Effective stellar temperature versus incident stellar flux (Seff) for the outer edge. The CO2
maximum greenhouse limit (dashed) is shown along with the empirical outer edge (solid black) and outer
edge limits containing 5%, 10%, 20%, 30%, and 50% H2 (red solid)
5
Figure 2: Effective stellar temperature versus (left) incident stellar flux (Seff) and (right) orbital distance
for the classical (dashed), empirical (solid black), and volcanic hydrogen (solid red) habitable zone. The
red curves contain 30% and 50% H2, respectively. As shown in Kopparapu et al. (2013), Earth appears
near the classical inner edge because of the generic model assumption of 100% relative humidity. With
subsaturation, Earth is well within the classical HZ (e.g., Leconte et al. 2013)
The volcanic hydrogen HZ boundaries can be
calculated using equation 1 with constants
from Table 1 and Table 2:
(1)
where T* is equal to Teff – 5780K and Seff(sun)
are the fluxes computed for our solar system
(see Tables 1-2). The corresponding orbital
distances of the HZ can be calculated using
equation 2 (fig.2):
(2)
where L/Lsun is the stellar luminosity in solar
units and d is the orbital distance in AU.
Planetary mass that is linked to planetary
model surface pressure via gravity and
planetary radius have only a small effect on
the HZ limits (Kopparapu et al. 2014).
6
Although the inner edge of the HZ is also
affected by the addition of H2, the runaway
greenhouse limit distance (e.g. Kasting et al.,
1993) only moves outward by ~0.1% to 4%
for 1% H2 and 50% H2 (fig. 2), respectively,
because of the dense water vapor atmosphere
at the inner edge. At temperatures above
~300 K water vapor absorption begins to
mask hydrogen collision-induced absorption,
decreasing its effectiveness.
The greenhouse effect of CO2
increases with higher pCO2 to a maximum
value (fig. 3), corresponding to the maximum
greenhouse effect. However, as the hydrogen
concentration increases, the CO2 partial
pressures associated with the outer edge
occur at
larger orbital
distances. This occurs because cooling from
both CO2 condensation and scattering
eventually outpace the combined greenhouse
effect of CO2 and H2 farther from the host
star.
lower Seff and
Seff=Seff(sun)+aT*+bT*2+cT*3+dT*4d=LLsunSeff
Figure 3: Dependence of the incident stellar flux (Seff) on atmospheric CO2 partial pressure (pCO2) for the
solar system outer edge. The pCO2 values corresponding to the maximum greenhouse for atmospheres with
hydrogen (red curves) and without (dashed curve) are shown as vertical lines for 0% to 50% H2 (see text).
to
is added
About 8 bars of CO2 are required to
sustain surface temperatures above freezing
at the CO2 maximum greenhouse limit of the
classical HZ in our solar system (Kasting et
al., 1993; Kopparapu et al., 2013). When
hydrogen
the planetary
atmosphere, the CO2 partial pressure required
at the outer edge of the volcanic-H2 HZ
decreases to 6 bar with 1% H2, and to 1 bar at
30% H2 (fig. 3). The corresponding H2 partial
pressures (with 1 bar N2) are 0.071 bar and
0.857 bar respectively. At 50% H2, the CO2
and H2 partial pressures at the outer edge of
the volcanic-hydrogen HZ are 0.5 and 1.5
bars, respectively.
The stellar energy distribution of host
stars also determines what CO2 concentration
7
the maximum CO2 greenhouse effect is
reached. For cooler stars, increased near-
infrared absorption and reduced scattering
lower this CO2 concentration (e.g. Kasting et
al., 1993). For a cool M6 star (Teff ~ 3000 K),
the CO2 partial pressure at the outer edge of
the classical HZ is 15 bar. Adding H2 reduces
the required amount of CO2 at the outer edge
to 9 bar and 1 bar for 1% and 30% H2,
respectively. The corresponding H2 partial
pressures are 0.1 bar and 0.857 bar.
Atmospheric scale height increases with
the addition of hydrogen. For example,
adding 30% H2 to the atmosphere of a
habitable planet located in our solar system's
outer edge (1 bar N2 and 8 bar CO2) increases
its atmospheric scale height by over a factor
a
assuming
of 1.5 due to the decreased atmospheric mean
molecular mass,
similar
temperature structure. In addition, such H2-
rich planets at the outer edge of the HZ also
have less CO2 in their atmospheres, reducing
the optical density as compared to those CO2
atmospheres that do not contain hydrogen.
These
the
detection of atmospheric spectral features
with the next generation of telescopes,
making such planets very interesting targets
for JWST and the E-ELTs in the search for
life.
two effects could facilitate
4. DISCUSSION
in
the mantles
Although the present day mantle of Earth
is relatively oxidized, it is generally assumed
that
terrestrial planets,
including Earth, start reduced (i.e. oxygen-
poor), which may lead to conditions that
promote high levels of H2 outgassing. One of
the main arguments for a reduced early
mantle is that Earth (and possibly other
terrestrial planets) must have entered a
magma ocean stage during accretion, which
would have acted as a surficial oxygen sink
(e.g., Scaillet
2011).
Afterwards, planetary mantles are thought to
become oxidized over time, either gradually
or in a stepwise fashion (Wood et al., 2006).
Mantle oxidation on Earth may have
occurred rather quickly, perhaps only ~100
million
core-mantle
differentiation (Trail et al., 2011).
and Gaillard,
years
after
Mantle oxidation may have taken longer
on Mars though. High H2 outgassing on early
Mars is inferred from Martian meteorites
suggesting a highly reduced ancient Martian
mantle (Stanley et al., 2011; Grott et al.,
2011). If early Mars was kept warm by
atmospheric H2 up until ~3.6 Ga, this would
suggest reducing mantle conditions had
persisted for ~1 billion years. Supporting this
view, geochemical results from the ~4 billion
8
year old meteorites, ALH84001 and NWA
Black Beauty, infer a reduced mantle for at
least ~0.5 billion years (e.g. Grott et al.,
2011).
These solar system examples warrant
some discussion of how mantle oxidation
may generally proceed on planets. One idea
suggests lower mantle pressures of larger
terrestrial bodies (e.g. Earth) are sufficiently
high to convert iron(II) oxide to iron metal
and iron(III) oxide, oxidizing the mantle
during accretion (Wade and Wood, 2005).
According to this notion, smaller planets like
Mars would not have generated internal
pressures sufficiently high
for mantle
oxidation, in agreement with the meteorite
evidence. This may suggest that atmospheric
H2 may not easily build up on more massive
planets. However, this may happen in a
number of ways. First, these calculations
assume that H2 escapes at the diffusion limit,
but H2 concentrations can potentially build
up if escape rates (including hydrodynamic
and Jean's) are lower than this limit, as had
been suggested for the early Earth (Tian et al.
2005). Hydrogen retention will also be
favored on planets with stronger magnetic
fields or on those with higher outgassing rates
per unit area.
Moreover, it is possible that H2- rich
atmospheres may be longer-lived on super-
Earths. First, their enhanced gravity results in
reduced H2 escape rates, promoting hydrogen
retention (e.g. Pierrehumbert and Gaidos,
2011). Secondly, mantle oxidation may be
slowed as a result, increasing the lifetime of
reduced mantle conditions. Larger planets
may have higher outgassing rates, possibly
leading to a feedback in which higher
outgassing rates enhance weathering rates
and vice versa (e.g. Heller 2015). Plus,
interior heat is retained longer on super-
Earths, ensuring that plate tectonics and
volcanism both last (ibid).
Hydrogen retention is also favored for
outer edge planets. Hamano et al. (2013)
suggest that selective hydrodynamic escape
of H over O during accretion could lead to
early mantle oxidation. However, outer edge
planets would be less susceptible to this
effect because corresponding EUV fluxes are
much lower than in the inner solar system.
Given that hydrodynamic escape rates are ∝
1/(d(AU)2), and assuming energy-limited
escape (e.g. Ramirez and Kaltenegger, 2014),
H escape rates (all else equal) on a young
accreting Earth-mass planet orbiting the early
Sun at 1.7 AU would have been ~ a factor of
~3 – 6 lower than on early Earth or Venus.
These loss rates would decrease further by
1/30 – 1/60 on a 10-Earth mass super-Earth.
It is also possible that accreting planets
incorporate more hydrogen because
the
protoplanetary disk itself is enriched in the
element. In addition, because stellar fluxes
are lower for these distant planets, magma
ocean durations should decrease as well,
further weakening escape (Hamano et al.
2013). All of these factors could more than
offset the factor of ~15 – 40 relatively higher
H2 outgassing rates inferred for a highly
reduced mantle (Ramirez et al., 2014). Thus,
sizeable hydrogen
larger
terrestrial planets located near the outer edge
may be sustainable as long as mantle
conditions support conditions that favor
hydrogen retention over escape to space.
inventories on
Certain atmospheric spectral features,
including N2O and NH3, which can, but do
not have to be produced biotically, could be
detected
in H2-dominated atmospheres
(Seager et al., 2013; Baines et al., 2014).
Such volcanic-hydrogen atmospheres may
also be able
to evolve methane-based
photosynthesis
al., 2014).
(Bains
Distinguishing biosignatures from abiotic
sources
in such atmospheres will be
challenging. NH3 can be formed abiotically
through
in
reaction of N2 and H2
et
9
hydrothermal vents on planets with reducing
mantles (Kasting et al., 2014). N2O can be
formed a number of ways including through
atmospheric shock from meteoritic fall-in,
lightning, UV radiation (e.g. Ramirez, 2016)
and through solar flare interactions with the
magnetosphere (Airapetian et al., 2016).
Thus, future biosignature studies should
focus on modeling biotic and abiotic sources
for these gases in thin volcanic-hydrogen
atmospheres.
Lastly, earlier studies suggesting that
backscattering from CO2 clouds could move
the outer edge beyond 2 AU (Mischna et al.,
2000) is challenged by new work indicating
that the greenhouse effect of CO2 clouds may
be greatly overstated (Kitzmann et al., 2016;
2017). Even should warming from CO2
clouds be relatively ineffective, volcanically
outgassed H2 in some stellar systems may
compensate and keep the HZ similarly wide
for geologically significant timescales.
5. CONCLUSION
We have calculated the limits of the
volcanic-H2 HZ
for H2 concentrations
ranging from 1% to 50%. We find that the
addition of 30% H2 can extend the outer edge
in our solar system to ~2.23 AU (2.4 AU with
50% H2). The orbital distance of the outer
edge of the volcanic-hydrogen HZ moves
outward by as much as ~60% as compared to
the classical HZ for A to M type main
sequence Stars. The inner edge only moves
out by ~0.1% to 4%.
Reduced mantle conditions that promote
hydrogen outgassing on small terrestrial
planets may be maintained
long
timescales (>0.5 Gyr). Atmospheric H2 may
be retained on larger terrestrial planets,
including
through higher
gravity and potentially stronger magnetic
fields. H retention is also generally favored in
super-Earths,
for
planets with higher outgassing rates, and if
H2 escapes more slowly than the diffusion
limit. Hydrodynamic escape rates are also
lower for planets farther away from the host
star, mitigating the selective escape of H over
O. Generally, hydrogen-rich atmospheres
composed of N2, CO2 and H2O can be
maintained as long as volcanic outgassing of
H2 outpaces its escape.
Moreover, the addition of H2 increases
the atmospheric scale height making super-
Earths near the volcanic-H2 HZ outer edge
interesting observational targets. Near-future
missions like JWST and the ELTs could
potentially detect atmospheric
features,
including biosignatures, on such planets.
Table 1. Coefficients to calculate the outer edge of the volcanic hydrogen HZ
Constant
1% H2
5% H2
10% H2
20% H2
30% H2
50% H2
Seff(Sun)
0.3129
0.2671
0.2452
0.2271
0.2009
0.1723
A
B
C
D
4.6597x10-05
3.4301x10-5
2.9373x10-5
2.5976x10-5
1.7821x10-5
1.4029x10-05
3.7621x10-10
1.2999x10-10
2.3132x10-10
-1.0703x10-9
-6.9474x10-11
-3.2318x10-10
-9.332x10-13
-6.6514x10-13
-5.1625x10-13
-6.8630x10-13
-3.1860x10-13
-2.8689x10-13
9.4571x10-17
6.3759x10-17
4.9119x10-17
1.1099x10-16
3.3181x10-17
4.2926x10-17
Table 2. Coefficients to calculate the inner edge of the volcanic hydrogen HZ
Constant
1% H2
5% H2
10% H2
20% H2
30% H2
50% H2
Seff(Sun)
0.9975
0.9946
0.9884
0.9775
0.9655
0.9377
a
b
c
d
1.0708x10-4
1.0644x10-4
1.0613x10-4
1.0501x10-4
1.0383x10-4
1.0123x10-4
8.5594x10-9
7.4458x10-9
8.4633x10-9
8.3633x10-9
8.2456x10-9
7.9833x10-9
-2.3600x10-12
-2.4365x10-12
-2.3376x10-12
-2.3106x10-12
-2.2837x10-12
-2.2204x10-12
1.2683x10-16
2.0284x10-16
1.2619x10-16
1.2447x10-16
1.2364x10-16
1.2098x10-16
10
ACKNOWLEDGEMENTS
The authors would like to thank Robin Wordsworth for kindly providing his CO2-H2 collision-
induced absorption data. We also thank James F. Kasting for the helpful discussions. The authors
acknowledge support by the Simons Foundation (SCOL # 290357, Kaltenegger) and the Carl
Sagan Institute.
REFERENCES
Airapetian, V. S., Glocer, A., Gronoff, G., Hébrard, E., & Danchi, W. 2016, NatGe., 9, 452- 455
Bains, W., Seager, S., & Zsom, A. (2014). Life, 4(4), 716-744.
Baranov, Y. I., Lafferty, W. J., & Fraser, G. T. 2004, Journal of molecular spectroscopy, 228,2,
432-440.
Batalha, N., Domagal-Goldman, S. D., Ramirez, R., & Kasting, J. F. 2015, Icarus, 258, 337-349.
Borysow, A. 2002, A&A, 390, 2, 779-782.
Grott, M., Morschhauser, A., Breuer, D., & Hauber, E. 2011, Earth and Planetary Science
Letters, 308, 3, 391-400.
Gruszka, M., & Borysow, A. 1997, Icarus, 129, 1, 172-177.
Gruszka, M. Borysow, A. Molecular physics 93.6 (1998): 1007-1016.
Haghighipour N. & Kaltenegger, L, Habitability of Binary Systems II: P-type binaries, ApJ, 777,
2, 166, 13 pp., 2013
Hamano, K., Abe, Y., & Genda, H. 2013, Natur, 497, 7451, 607-610.
Heller, René, 2015. Scientific American 312,1, 32-39.
Kaltenegger, L. & Haghighipour N., Habitability of Binary Systems I: S-type binaries, ApJ, 777,
2, 165, 11 pp., 2013
Kaltenegger, L., Selsis, F., Fridlund, M. et al. (2010). AsBio, 10(1), 89-102
Kasting, J. F., Kopparapu, R., Ramirez, R. M., & Harman, C. E. 2014, PNAS, 111, 35, 12641-
12646.
Kasting, J.F., Whitmire, D.P., Reynolds, R.T., 1993, Icarus 101 108-128
Kitzmann, D. 2016, ApJL, 817, 2, L18.
11
Kitzmann, D. (2017). arXiv preprint arXiv:1701.07513.
Kopparapu, R. K., Ramirez, R., Kasting, J. F., Eymet, V., Robinson, T. D., Mahadevan, S., ... &
Deshpande, R. 2013, ApJ, 765, 2, 131.
Kopparapu, R. K., Ramirez, R. M., SchottelKotte, J., Kasting, J. F., Domagal-Goldman, S., &
Eymet, V. 2014, ApJL ,787, 2, L29.
Leconte, J., Forget, F., Charnay, B., Wordsworth, R., & Pottier, A. (2013). Nature, 504(7479),
268-271.
Mischna, M. A., Kasting, J. F., Pavlov, A., & Freedman, R. 2000, Icarus, 145, 2, 546-554.
Montési, L. G., & Zuber, M. T. (2003). JGR:Planets, 108(E6).
Pierrehumbert, R., & Gaidos, E. 2011, ApJL, 734, 1, L13.
Ramirez, R. M., Kopparapu, R., Zugger, M. E., Robinson, T. D., Freedman, R., & Kasting, J. F.
2014, NatGe, 7, 1, 59-63.
Ramirez, R. M., & Kaltenegger, L. 2014, ApJL, 797, 2, L25.
Ramirez, R. M., & Kaltenegger, L. 2016, ApJ, 823, 1, 6.
Ramirez, R.M. 2016, NatGe,9, 6, 413-414.
Scaillet, B., & Gaillard, F. 2011, Nature, 480, 7375, 48-49.
Seager, S., Bains, W., & Hu, R. 2013, ApJ, 777, 2, 95.
Stanley, B. D., Hirschmann, M. M., & Withers, A. C. 2011, Geochimica et Cosmochimica
Acta, 75, 20, 5987-6003.
Tian, F., Toon, O. B., Pavlov, A. A., & De Sterck, H. 2005, Science, 308, 5724, 1014-1017.
Trail, D., Watson, E. B., & Tailby, N. D. 2011, Natur, 480, 7375, 79-82.
Stevenson, D. J. 1999, Natur, 400, 6739, 32-32.
Wade, J., & Wood, B. J. 2005, E&PSL, 236, 1, 78-95.
12
Walker, J. C. 1977, New York: Macmillan, and London: Collier Macmillan, 1977, 318 pp.
Wood, B. J., Walter, M. J., & Wade, J. 2006, Natur, 441, 7095, 825-833.
Wordsworth, R. 2012, Icarus, 219, 1, 267-273.
Wordsworth, R., & Pierrehumbert, R. 2013, Science, 339, 6115, 64-67.
Wordsworth, R., Kalugina, Y., Lokshtanov, S., et al. (2017). GeoRL. doi:
10.1002/2016GL071766
Zsom, A., Kaltenegger, L., & Goldblatt, C. (2012). Icarus, 221(2), 603-616.
13
|
0902.3999 | 1 | 0902 | 2009-02-23T21:01:14 | The Two Modes of Gas Giant Planet Formation | [
"astro-ph.EP"
] | I argue for two modes of gas giant planet formation and discuss the conditions under which each mode operates. Gas giant planets at disk radii $r>100$ AU are likely to form in situ by disk instability, while core accretion plus gas capture remains the dominant formation mechanism for $r<100$ AU. During the mass accretion phase, mass loading can push disks toward fragmentation conditions at large $r$. Massive, extended disks can fragment into clumps of a few to tens of Jupiter masses. This is confirmed by radiation hydrodynamics simulations. The two modes of gas giant formation should lead to a bimodal distribution of gas giant semi-major axes. Because core accretion is expected to be less efficient in low-metallicity systems, the ratio of gas giants at large $r$ to planets at small $r$ should increase with decreasing metallicity. | astro-ph.EP | astro-ph |
The Two Modes of Gas Giant Planet Formation
Aaron C. Boley
Institute for Theoretical Physics, University of Zurich, Winterthurerstrasse 190, Zurich,
CH-8057, Switzerland; [email protected]
ABSTRACT
I argue for two modes of gas giant planet formation and discuss the conditions
under which each mode operates. Gas giant planets at disk radii r > 100 AU
are likely to form in situ by disk instability, while core accretion plus gas capture
remains the dominant formation mechanism for r < 100 AU. During the mass
accretion phase, mass loading can push disks toward fragmentation conditions
at large r. Massive, extended disks can fragment into clumps of a few to tens of
Jupiter masses. This is confirmed by radiation hydrodynamics simulations. The
two modes of gas giant formation should lead to a bimodal distribution of gas
giant semi-major axes. Because core accretion is expected to be less efficient in
low-metallicity systems, the ratio of gas giants at large r to planets at small r
should increase with decreasing metallicity.
Subject headings: planetary systems:
planetary disks — hydrodynamics — instabilities — radiative transfer
formation — planetary systems: proto-
1.
Introduction
Both core accretion plus gas capture (e.g., Pollack et al. 1996) and direct formation
by disk instability (Cameron 1978; Boss 1997) can form gas giant planets in principle. In
order for gravitational instabilities (GIs) to form gas giants directly, the Toomre (1964)
Q = csκ/πGΣ, where cs is the sound speed, κ is the epicyclic frequency, and Σ is the surface
density, must approach unity and the local cooling time must be (cid:46) the local orbit period
(Gammie 2001; Rice et al. 2005; Rafikov 2005, 2007). An isothermal, low-Q is strongly
susceptible to fragmentation (e.g., Tomley et al. 1994; Boss 1998; Nelson et al. 1998; Mayer
et al. 2002; Pickett et al. 2003).
Analytical work (e.g., Rafikov 2007) and radiation hydrodynamics simulations (Nelson
et al. 2000; Cai et al. 2006, 2008; Boley et al. 2006, 2007b; Stamatellos & Whitworth 2008a;
Boley & Durisen 2008) find that disk fragmentation inside r ∼ tens of AU is unlikely because
– 2 –
regions with high cooling rates and low Q are rare (cf, most recently, Boss 2008 and Mayer et
al. 2007). When Q is low, the gas is cool, so cooling times are typically long. When cooling
times are short, the gas is hot, so Q is typically high. In contrast, Boss (2004) argues that
convection can increase the cooling rates enough to cause fragmentation. Likewise, Mayer
et al. (2007) report convection-induced fragmentation in their simulations when the mean
molecular weight is increased from 2.4 to 2.71. Boley et al. (2006) and Boley & Durisen
(2008) argue in return that the reported convective motions are likely shock bores along the
spiral shocks (Boley & Durisen 2006) or an artifact of suddenly changing the mean molecular
weight. Moreover, Rafikov (2007) and Boley et al. (2006, 2007b) find that convection does
not decrease cooling times enough to trigger fragmentation because sustained convection is
regulated by the entropy gradient and energy must be radiated away at the disk’s surface.
Core accretion seems to be the only option for gas giant formation inside r ∼ 100 AU.
However, the recently-discovered substellar companions with wide semi-major axes a ∼ 100
AU and masses M ∼ few to tens MJ (e.g., Luhman et al. 2006; Lafreniere et al. 2008; Kalas
et al. 2008) challenge the interpretation that core accretion is the sole formation mechanism
for gas giants. The susceptibility of disks to fragmentation at large r remains an open
question. Stamatellos et al.
(2007; hereafter SHW2007) and Stamatellos & Whitworth
(2009; hereafter SW2009) find that extended, massive disks can fragment under realistic
conditions (cf Boss 2006); but, their results are based on running simulations with highly
unstable initial conditions, which as argued here, can significantly change the outcome of
disk evolution.
of the flux, where ∇ · F = 4ρκpσT 4 for low optical depth τ =(cid:82) ∞
Consider a disk where the sound speed becomes constant with radius for large r. For
Σ ∼ r−q, Q ∼ rqΩ ∼ rq−1.5. As long as q < 1.5, the stabilizing shear contribution to Q falls
off faster than the destabilizing self-gravity contribution, and the disk can become susceptible
to GIs. Now consider the radiative cooling time for some fluid element trad = /∇·F, where
is the internal energy density of the gas. Using the free-streaming limit for the divergence
−∞ ρκP dz, the cooling time
can be approximated by trad ∼ Σc2
s/(4τ σT 4), where ρ is the mass density at temperature
T and κP is the Planck mean opacity. For τ (cid:28) 1, radiative cooling becomes extremely
inefficient. Because cooling in the optically thick disk is also inefficient, the τ ∼ 1 transition
region may be the most conducive regime for disk fragmentation (e.g., Rafikov 2005). The
cooling time in a disk with Σ ∼ 1 g cm−2, τ ∼ 0.1, and T ∼ 10 K, is trad ∼ 600 yr, which is
less than the orbital period around a 1 M(cid:12) star at r ∼ 100 AU.
What if the τ ∼ 1 and Q ∼ 1 regions do not coincide? Mass loading, i.e., mass accretion
1Boley et al. (2006) simulated a disk with a µ ∼ 2.7; they did not see fragmentation.
– 3 –
onto the disk, is crucial for driving fragmentation at r > 100 AU. Numerical fragmentation
studies have focused mainly on thermal energy balance arguments. However, there are
additional considerations when evaluating the stability of a disk against fragmentation. The
rate that Q changes in a local region of a disk is
d ln Q
dt
=
1
2
d ln c2
s
dt
+
d ln κ
dt
− d ln Σ
dt
.
(1)
When GIs are in a self-regulating phase, the effective Shakura & Sunyeaev (1973) α tends
toward a few hundredths or less (e.g., Gammie 2001; Lodato & Rice 2004; Mejia et al. 2005;
Boley et al. 2006), so the timescale for Σ and κ to change due to mass transport should be
much larger than the local dynamical time. For fragmentation criteria studies of isolated
disks, the thermal energy term is the most relevant. In order to avoid fragmentation, the
local cooling time for the gas tcool > f (γ)td, where td is the local gas orbital period and f (γ)
is some function of the adiabatic index of order unity. In the local model formalism, the tcool
criterion is associated with a critical effective αc ∼ 0.06 (Gammie 2001; Rice et al. 2005).
This α-limit indicates the point at which local energy dissipation by GIs cannot balance
cooling, and as a result, the disk fragments.
Mc ≈ 3αcc3
In addition to energy transport, αc can be related to a critical mass flux Mc ≈ 3παcc2
sΣ/Ω.
s/(GQ) (see also Goodman 2003; Matzner
Substituting Q into this relation gives
& Levin 2005). At large r, where the temperature of the disk approaches the envelope tem-
perature, the ratio Mc/ Me ∼ 3αc/Q, where Me ∼ c3
s/G is the mass accretion onto the disk
Me ∼ 2 × 10−5M(cid:12) yr−1.
from the envelope. For an envelope temperature of T = 30 K,
Because αc ∼ 0.06, mass loading can operate at multiple times the maximum local transport
flux. Episodes of non-local transport and/or disk fragmentation should occur for realistic
envelope accretion rates.
I present results from radiative hydrodynamics simulations that show mass loading can
drive a disk toward fragmentation at large r. The fragments form clumps with masses greater
than a few MJ . Throughout this Letter, I refer to the inner disk as the region inside r ∼ 100
AU and the extended disk as the region outside this boundary. This transition radius is
illustrative, but represents a reasonable estimate for where fragmentation becomes possible
(e.g, Matzner & Levin 2005).
2. Methodology
I use CHYMERA (Boley 2007) to model the formation of massive extended disks.
CHYMERA is an Eulerian code that solves for the equations of hydrodynamics with self-
gravity on an evenly spaced, cylindrical grid. The rotational states of molecular hydrogen
– 4 –
are taken into account (Boley et al. 2007a), and the radiative transfer scheme has passed a
series of analytic tests. The radiative transfer algorithm and test results are presented in
Boley et al. (2007b). D’Alessio et al. (2001) opacity tables are used for calculating optical
depths, with a maximum grain radius amax = 1µm. In the following simulations, an equilib-
rium ortho:parahydrogen ratio (O:P) is used. The gas is assumed to have a mixture of 0.73,
0.25, and 0.02 for hydrogen, helium, and metals, respectively.
Preliminary simulations using the CHYMERA radiative transfer scheme show that the
radiative cooling time is shorter than the local orbital time and that the disk is effectively
isothermal, in agreement with the estimate in §1. For modeling spiral waves, the normal
radiative transfer scheme is stable in these simulations because the gas temperature is never
far from the incident irradiation temperature. However, for some preliminary simulations,
the cooling algorithm became numerically unstable immediately after clump formation due
to the sudden increase in temperature, creating highly disparate radiative transfer and hy-
drodynamics timescales. In order to study fragmentation at large r, the following radiative
cooling algorithm is used for the simulations presented here. The divergence of the flux
∇ · F = −(A/V )σ(T 4 − T 4
irr)f−1
τ
,
(2)
where A/V is the cell area-to-volume ratio, Tirr is the incident irradiation on the disk at r, and
fτ = ∆τ + 1/∆τ . The local optical depth across a cell is calculated by ∆τ = ρ(κRosseland(1−
exp[−2∆τPlanck]) + κPlanck exp[−2∆τPlanck])V 1/3, where ∆τPlanck is an initial estimate using
the Planck mean opacity. This cooling approximation goes to the free-streaming limit for
small ∆τ and to zero for large ∆τ . In order to ensure that the algorithm is stable for large
hydrodyamic time steps, the divergence of the flux is adjusted such that
∇ · F adjusted = ∇ · F exp[−(∆thydo/∆trad)2]
+ ρ(eequil − e)/∆thydro(1 − exp[−(∆thydro/∆trad)2]).
(3)
Here, e is the specific internal energy of the gas, eequil is the internal energy of the gas if
it were at Tirr, ∆thydro is the Courant time step, and ∆trad = ρe/∇ · F is the radiative
timescale. The irradiation temperature for these simulations is set to 30 K for all r.
The central protostar’s position is integrated self-consistently with a softened potential,
where Φstar = GMstar/(r− r(cid:48)2 + s2)1/2 for softening parameter s. In order to treat the force
on the star, the mass in each cell is treated as a point mass at the cell’s center, with the same
softening parameter used for the star. The star’s position is integrated from step i to i + 1
by the following: vi = vi−1/2 + 0.5ai∆ti−1, vi+1/2 = vi + 0.5ai∆ti, and xi+1 = xi + vi+1/2∆ti.
This algorithm is sufficient for maintaining, on average, the system’s center of mass at the
grid center.
– 5 –
Each disk is evolved on an r, φ , z = 256, 512, 64 grid, with a spatial resolution of ∆r,
r∆φ, and ∆z = 2, 2πr/512, 2 AU. Mirror symmetry is assumed about the midplane, and the
outer grid boundaries are outflow boundaries. There is an outflow boundary near the star,
but negligible mass passes through it in these simulations. Mass is added to the grid near
the top of the z outflow boundary between r = 60 and 300 AU. The added mass is given a
density profile of ρ ∼ r−p, a specific angular momentum (GMstarr3/(r2 + s2))1/2, an initial
vr = 0, and an initial vz = −(2GMstar/r)1/2. Upon reaching a disk mass of 0.1 M(cid:12), a random
density perturbation is imposed with a maximum variation of ±10%. For all simulations, the
disk is stable against GIs when the noise is added. For two of the simulations (see below) a
softening of s = 20 AU was applied. This softening causes an error in the epicyclic frequency
of just under 10% at r ∼ 50 AU when compared with the Keplerian frequency and 3% at
r ∼ 100 AU.
3. Simulations and Results
Four simulations (SIMA, SIMB, SIMC, and SIMD) are shown in Figure 1. All disks
except for SIMC fragment, and all disks evolve isothermally except in very high-density
regions. Because the vertically integrated midplane τ never becomes larger than a few,
except in clumps, the above radiative transfer approximation is reasonable. Clumps reach
temperatures in excess of 100 K.
(1) SIMA: The protostar is set to 0.3 M(cid:12). The initial mass loading Md ∼ 10−4M(cid:12) yr−1
until the disk mass Md = 0.1M(cid:12), after which Md is reduced to 10−5M(cid:12) yr−1. The infall
density profile is set to p = 1.5. A softening s = 20 AU is used because the accretion rate is
less than what is used in the other simulations, requiring a longer evolution. The simulation
is evolved for 16400 yr (3 P200, orbital periods at r ∼ 200 AU), and reaches a disk mass of
0.21 M(cid:12) before the simulation is stopped. Dense spiral waves develop, and a condensation
forms near r ∼ 70 AU. Fragmentation appears to be near wave corotation, in agreement
with Durisen et al. (2008). By the end of the simulation, the clump has grown to 20 MJ
and is located at r ∼ 110 AU. As a cautionary check, a portion of this simulation was rerun
with CHYMERA’s normal radiative transfer algorithm. The disk behaved isothermally.
(2) SIMB: The protostar is set to 1 M(cid:12), and Md ∼ 10−4M(cid:12) yr−1 for the duration of the
simulation. The simulation is evolved for 5000 yr (∼ 1.8 P200), and the disk grows to 0.52
M(cid:12). The infall density profile p = 1.5, and no softening is applied. Mass loading drives Q
below unity outside r ∼ 100 AU, and the disk fragments into 11 condensations by the time
the simulation is stopped, ranging in mass from ∼ 4 to 14 MJ .
– 6 –
(3) SIMC: Similar to SIMB, but the mass accretion is halted when Md = 0.33M(cid:12). At
this mass, Q in a region near r ∼ 90 AU drops below unity. This simulation investigates
whether the disk can recover from the initial mass loading and avoid fragmentation. When
GIs did set in, they rapidly transported mass both away and toward the star, so a softening
of s = 20 AU was applied and the inner disk boundary was moved from r ∼ 20 AU to r ∼ 6
AU. The initial simulation without softening shows a similar behavior, but was not evolved
beyond the initial burst of GIs. The burst transports mass efficiently and Q reaches a mass-
weighted average Q ∼ 1.3 between r ∼ 100 and 200 AU. This simulation is evolved for 9200
yr (∼ 3.3 P200). Although a longer integration may be required to ascertain whether the disk
will eventually fragment, the simulation does indicate that fragmentation is not guaranteed
in the outer disk.
(4) SIMD: The protostar is set to 1 M(cid:12), and the mass accretion rate is 10−4M(cid:12) yr−1.
As in SIMC, the mass accretion is halted once Md = 0.33M(cid:12). The density profile for the
infalling material p = 0.5. This choice for p places the minimum Q further out in the disk,
such that Q first drops below unity near r ∼ 140 AU. Unlike SIMC, the spiral waves are
unable to redistribute the mass enough to avoid fragmentation. The disk forms a 6 MJ
clump at r ∼ 140 AU, which is subsequently transported to just inside r ∼ 100 AU and
grows to 11 MJ by the end of the simulation. The disk is evolved for 8600 yr (∼ 3 P200).
Wide semi-major axis, a, gas giants (WaGGs) and at least one brown dwarf (BD) are
formed in these simulations. All disks go through strong burst-like phases, consistent with the
arguments in §1. Although a high M is used for SIMB, SIMC, and SIMD, disk fragmentation
M ∼ 10−5M(cid:12) yr−1, as seen in SIMA. The local Jeans length is resolved
is still expected for
by at least four cells throughout the simulations (Truelove et al. 1997; Nelson 2006), where
the local cell size is taken to be the geometric mean of the cell dimensions. These results
indicate that fragmentation at large disk radii should be common and that WaGGs can be
explained by in situ formation. These simulations demonstrate that the number or fragments
and the fragment masses depend on how the disk is assembled. Studies investigating the
clump mass spectrum of extended, massive protoplanetary disks, e.g., SW2009, are very
sensitive to initial conditions.
Although the fate of these clumps is unknown, the results of Mayer et al. (2004) and
SW2009 indicate that some condensations can survive. Radiative feedback, which is ne-
glected in the current radiative cooling algorithm, should limit clump growth. Detailed
simulations that can resolve the photosphere of a fragment are required to address this
concern, which is beyond the resolution limit in these simulations.
These disks are evolved through only one burst. Vorobyov & Basu (2005; hereafter
VB2005) found in their 2D simulations that several bursts can occur during the disk accretion
– 7 –
phase. Whether a system forms WaGGs/BDs and retains them may be dependent on the
details of the last GI burst. Because these simulations suggest that disks should be susceptible
to fragmentation during their formation, they also support the possibility of a clump-driven
outburst mechanism for FU Orionis objects (VB2005). The FU Orionis phenomenon is
characterized by a rapid increase in the optical brightness of a young T Tauri object, typically
5 magnitudes over a few to tens of years. Accretion from the inner disk onto the star
is estimated to be as high as 10−4M(cid:12) yr−1 (Hartmann & Kenyon 1996). Because FU Ori
objects have decay timescales of ∼ 100 yr, an entire minimum mass solar nebula (∼ 0.01 M(cid:12))
can be accreted onto the protostar during an event. The best explanation for the optical
outburst is a thermal instability (TI) (Bell & Lin 1994). However, prodigious mass flux
through r ∼ 0.1 AU seems to be required to drive the TI. A tidally-disrupted WaGG could
supply the inner disk with 0.01 M(cid:12) mass, even when only the extended disk is gravitationally
unstable. Helled et al. (2006) calculated the contraction time for a Jupiter-mass clump, and
found that such a clump takes 3 × 105 yr to reach a central temperature of 2000 K, i.e.,
the temperature required to dissociate enough H2 to lead to rapid collapse. This suggests
that a WaGG has sufficient time to be transported inside r ∼ 1 AU before it reaches
mean densities that are high enough to avoid tidal disruption, where the disruption radius
rt ∼ 8 × 10−3(¯ρ/g cm−3)−1/3(M/M(cid:12))1/3 AU. The Helled et al. model remains at a mean
density of ¯ρ ∼ 2× 10−7 g cm−3 for ∼ 105 yr. This density corresponds to a disruption radius
rt ∼ 1 AU for a 0.3 M(cid:12) star. An immediate observational signature of this mechanism would
be a large shift in the radial velocity of the protostar just before an outburst.
4. The Two Modes of Planet Formation
Boss’s (1997) advancement of the disk instability mechanism has spawned a decade’s
worth of work examining GIs as a formation mechanism for gas giant planets (see Durisen
et al.’s 2007 review). One of the principal reasons for the mechanism remaining contested
for so long is that, like core accretion, the mechanism works under the right conditions.
Rafikov (2005, 2007), Boley et al. (2006, 2007b), Stamatellos & Whitworth 2008, and Boley
& Durisen (2008) argue that radiative cooling timescales are too long for fragmentation out
to r ∼ 40 AU and that sustained convection does not cause fragmentation. This is also
consistent with Nelson et al.’s (2000) 2D radiative hydrodynamics simulations, who assumed
a polytropic vertical density structure in order to calculate a photosphere temperature. Con-
vection pushes the entropy gradient toward zero, so a vertical polytropic density stratification
assumes efficient convection. Observationally, the planet-metallicity relationship (Valenti &
Fischer 2005) indicates that planet formation favors a high-metallicity environment, which is
strong evidence for core accretion (cf Boss 2005). The estimated core masses of Jupiter and
– 8 –
Saturn (Saumon & Guillot 2004; Militzer et al. 2008), along with the ice giants, support the
core accretion mechanism. This evidence suggests that the dominant formation mechanism
for gas giants inside r ∼ 100 AU is core accretion.
In contrast to the conditions inside r ∼ 100 AU, optical depths should approach unity
for a substantial ∆r in the extended disk, as occurs in these simulations. Efficient radiative
cooling, long orbital periods, and an equilibrium O:P ratio combine to favor fragmentation
in the extended disk. WaGG formation as a result of mass loading represents the first mode
of planet formation, and takes place in the first 105 yr of the disk’s lifetime. Core accretion,
which can continue after the main disk formation phase, represents the slower, second mode
of gas giant formation. The survival of WaGGs will likely depend on the disk and mass
accretion conditions during the last burst of GI activity.
Both core accretion and disk instability are able to form gas giants under suitable
In the inner disk, conditions favor core accretion, and in the extended disk,
conditions.
conditions favor disk instability. Scattering and planet-disk interactions should wash out
any strict desert between the two formation regimes, but a bimodal population of gas giant
planet semi-major axes should still be present. Because core accretion is expected to be less
efficient in low-metallicity systems, the ratio of WaGGs to planets at small r should increase
with decreasing metallicity.
I would like to thank the referee for pointing out a complementary study by Rafikov
(2009, astro-ph/0901.4739), which was posted during the review of this manuscript. Rafikov’s
work discusses fragmentation conditions and the stability of disks against high M ’s, while
the study presented here demonstrates that a disk can be pushed realistically toward frag-
mentation.
I thank R. H. Durisen, L. Mayer, G. Lake, R. Teyssier, O. Agertz, D. Stamatellos, and
the referee for comments that improved this manuscript. This research was supported by a
Swiss Federal Grant and the University of Zurich. The presented simulations were run on
NASA Advanced Supercomputing facilities.
– 9 –
REFERENCES
Bell, K. R., & Lin, D. N. C. 1994, ApJ, 427, 987
Boley, A. C. 2007, Ph.D. Thesis, Indiana University
Boley, A. C., & Durisen, R. H. 2006, ApJ, 641, 534
–. 2008, ApJ, 685, 1193
Boley, A. C., Durisen, R. H., Nordlund, A, & Lord, J. 2007b, ApJ, 665, 1254
Boley, A. C., Hartquist, T. W., Durisen, R. H., & Michael, S. 2007a, ApJ, 656, L89
Boley, A. C., Mej´ıa, A. C., Durisen, R. H., Cai, K., Pickett, M. K., & D’Alessio, P. 2006,
651, 517
Boss, A. P. 1997, Sci, 276, 5320, 1836
–. 1998, ApJ, 503, 923
–. 2004, ApJ, 610, 456
–. 2005, ApJ, 629, 535
–. 2006, ApJ, 637, L137
–. 2008, ApJ, 677, 607
Cai, K., Durisen, R. H., Michael, S., Boley, A. C., Mej´ıa, A. C., Pickett, M. K., & D’Alessio,
P. 2006, ApJ, 636, L149
Cameron, A. G. .W. 1978, M&P, 18, 5
D’Alessio, P., Calvet, N., & Hartmann, L. 2001, ApJ, 553, 321
Durisen, R. H., Boss, A., Mayer, L., Nelson, A., Quinn, T., & Rice, K. 2007, in Protostars
and Planets V, ed. B. Reipurth, D. Jewitt, & K. Keil (Tucson: Univ. Arizona Press),
607
Durisen, R. H., Hartquist, T. W., & Pickett, M. K. 2008, Ap&SS, 317, 3D
–. 2001, ApJ, 553, 174
Goodman, J. 2003, MNRAS, 339, 937
– 10 –
Hartmann, L., & Kenyon, S. J. 1996, ARA&A, 34, 207
Helled, R., Podolak, M., & Kovetz, A. 2006, Icarus, 185, Issue 1, 64
Kalas, P., Graham, J. R., Chiang, E., Fitzgerald, M. P., Clampin, M., Kite, E. S., Stapelfeldt,
K., Marois, C., Krist, J. 2008, arXiv:0811.1994
Lafreniere, D., Jayawardhana, R., & van Kerkwijk, M. H. (2008), ApJ, 689, L153
Lodato, G., & Rice, W. K. M. 2004, MNRAS, 351, 630
Luhman, K. L., Wilson, J. C., Brandner, W., Skrutskie, M. F., Nelson, M. J., Smith, J. D.,
Peterson, D. E., Cushing, M. C., & Young, E. 2006, ApJL, 649, 894
Matzner, C. D., & Levin, Y. 2005, ApJ, 628, 817
Mayer, L., Lufkin, G., Quinn, T., & Wadsley, J. 2007, ApJ, 661, 77
Mayer, L., Lufkin, G., Quinn, T., & Wadsley, J. 2002, Sci, 298, 1756
Mayer, L., Quinn, T., Wadsley, J., & Stadel, J. 2004, ApJ, 609, 1045
Militzer, B., Hubbard, W. B., Vorberger, J., Tamblyn, I., & Bonev, S. A. 2008, ApJ, 688,
L45
Mej´ıa, A. C., Durisen, R. H., Pickett, M. K., & Cai, K. 2005, ApJ, 619, 1098
Nelson, A. F. 2006, MNRAS, 373, 1039
Nelson, A. F., Benz, W., Adams, F. C., & Arnett, D. 1998, ApJ, 502, 342
Nelson, A. F., Benz, W., & Ruzmaikina, T. V. 2000, ApJ, 529, 357
Pickett, B. K., Mej´ıa, A. C., Durisen, R. H., Cassen, P. M., Berry, D. K., & Link, R. P. 2003,
ApJ, 590, 1060
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig,
Y. 1996, Icarus, 124, 63
Rafikov, R. R. 2005, ApJ, 621, L69
–. 2007, ApJ, 662, 642
Rice, W. K. M, Lodato, G., & Armitage, P. J. 2005, MNRAS, 364, L56
Saumon, D., & Guillot, T. 2004, ApJ, 609, 1170
– 11 –
Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337
Stamatellos, D., Hubber, D. A., & Whitworth, A. P. 2007, MNRAS, 382, L30
Stamatellos, D., & Whitworth, A. P. 2008, A&A, 480, 879
Stamatellos, D., & Whitworth, A. P. 2009, MNRAS, 392, 413
Tomley, L., Steiman-Cameron, T. Y., Cassen, P. 1994, ApJ, 422, 850
Toomre, A. 1964, ApJ, 139, 1217
Truelove, J. K., Klein, R. I., McKee, C. F., Holliman, J. H., II, Howell, L. H., Greenough,
J. A. 1997, ApJ, 489, L179
Valenti, J. A., & Fischer, D. A. 2005, ApJS, 159, 141
Vorobyov, E. I., & Basu, S. 2005, ApJ, 633, 137
This preprint was prepared with the AAS LATEX macros v5.2.
– 12 –
Fig. 1.— Surface density snapshots at the end of each simulation. Strong bursts of GI activity
occur in each disk. For three of the simulations, these bursts lead to fragmentation and clump
formation. Movies of these simulations can be viewed at http://www.aaroncboley.net under
the Movies tab.
|
1712.00437 | 1 | 1712 | 2017-12-01T18:38:41 | Tumbling motion of 1I/'Oumuamua reveals body's violent past | [
"astro-ph.EP"
] | Models of the Solar System evolution show that almost all the primitive material leftover from the formation of the planets was ejected to the interstellar space as a result of dynamical instabilities. Accordingly, minor bodies should also be ejected from other planetary systems and should be abundant in the interstellar space. The number density of such objects, and prospects for their detection as they penetrate through the Solar System, were speculated about for decades, recently rising high hopes with the Pan-STARRS and LSST surveys. These expectations materialized on 18 October 2017 with the Pan-STARRS's discovery of 1I/'Oumuamua. Here we report homogeneous photometric observations of this body from Gemini North, which densely cover a total of 8 hr over two nights. A combined ultra-deep image of 1I/'Oumuamua shows no signs of cometary activity, confirming the results from earlier, less sensitive searches. Our data also show an enormous range of brightness variations > 2.5 mag, larger than ever observed in the population of Solar System objects, suggesting a very elongated shape of the body. But most significantly, the light curve does not repeat exactly from one rotation cycle to another and its double-peaked periodicity of 7.5483 $\pm$ 0.0073 hr from our data is inconsistent with earlier determinations. These are clear signs of a tumbling motion, a remarkable characteristic of 1I/'Oumuamua's rotation, consistent with a catastrophic collision in the distant past. This first example of an impacted minor body of exosolar origin indicates that collisional evolution of minor body populations in other planetary systems is not uncommon. | astro-ph.EP | astro-ph | Tumbling motion of 1I/'Oumuamua reveals body's violent past*
M. Drahus1*, P. Guzik1, W. Waniak1, B. Handzlik1, S. Kurowski1, S. Xu2
1 Jagiellonian University, Kraków, Poland ([email protected])
2 Gemini Observatory, Hilo, HI, USA
Models of the Solar System evolution show that almost all the primitive material leftover from
the formation of the planets was ejected to the interstellar space as a result of dynamical
instabilities1. Accordingly, minor bodies should also be ejected from other planetary systems
and should be abundant in the interstellar space2. The number density of such objects, and
prospects for their detection as they penetrate through the Solar System, were speculated
about for decades3,4, recently rising high hopes with the Pan-STARRS5,6 and LSST7 surveys.
These expectations materialized on 18 October 2017 with the Pan-STARRS's discovery of
1I/'Oumuamua8. Here we report homogeneous photometric observations of this body from
Gemini North, which densely cover a total of 8 hr over two nights. A combined ultra-deep
image of 1I/'Oumuamua shows no signs of cometary activity, confirming the results from
earlier, less sensitive searches9–12. Our data also show an enormous range of brightness
variations > 2.5 mag9, larger than ever observed in the population of Solar System objects,
suggesting a very elongated shape of the body. But most significantly, the light curve does not
repeat exactly from one rotation cycle to another and its double-peaked periodicity of
7.5483±0.0073 hr from our data is inconsistent with earlier determinations9,10,13–15. These are
clear signs of a tumbling motion, a remarkable characteristic of 1I/'Oumuamua's rotation,
consistent with a catastrophic collision in the distant past. This first example of an impacted
minor body of exosolar origin indicates that collisional evolution of minor body populations in
other planetary systems is not uncommon.
* Based on observations obtained at the Gemini Observatory, which is operated by the Association of Universities for Research in Astronomy, Inc.,
under a cooperative agreement with the NSF on behalf of the Gemini partnership: the National Science Foundation (United States), the National
Research Council (Canada), CONICYT (Chile), Ministerio de Ciencia, Tecnología e Innovación Productiva (Argentina), and Ministério da Ciência,
Tecnologia e Inovação (Brazil).
1
1I/'Oumuamua8 is a newly discovered minor body on a strongly hyperbolic trajectory, which starkly
contrasts with even the most extreme cometary orbits bound to the Sun (Figure 1) and can only be
explained by having origin far in the interstellar space8,16,17. Unsurprisingly, the Galactic velocity
vector of this object is consistent with the median velocity vector of nearby stars17. 1I/'Oumuamua
entered the Solar system from the direction of the Solar apex with a large hyperbolic excess speed of
26 km s-1, then reached perihelion at 0.26 au from the Sun on 9 September 2017, approached the
Earth to 0.16 au on 16 October 2017, and is currently on its way back to the interstellar space.
Previous studies have not detected any signs of cometary activity9–12, found optical colors consistent
with Solar System's C-type and D-type asteroids10,13,14 and reported an average V-band absolute
magnitude of 22.8 to 23.010,12. Such an intrinsically faint body could be discovered only thanks to the
close Earth flyby in October 2017, which created a tremendous opportunity for the first detailed
characterization of a minor object of exosolar origin.
We were awarded Director's Discretionary Time on the 8.1 m Gemini North telescope (program GN-
2017B-DD-7) to observe this unique object. On 27 and 28 October 2017 we obtained a total of 442
r'-band18 images using a 30 sec integration time. The sky was photometric and median seeing close
to 0.6 arcsec in FWHM throughout the observations. We took data with the Gemini Multi-Object
Spectrograph (GMOS), which consists of three adjacent Hamamatsu CCDs, providing imaging over
a 5.5×5.5 arcmin field of view19. The instrument was configured to provide 2×2 binning and 0.1614
arcsec effective pixels.
The images were corrected for overscan, bias, and flatfield in the standard manner. Then, using our
established technique20, we subtracted background stars and galaxies interfering with 1I/'Oumuamua
as they moved across the field of view. As a result, we obtained uncontaminated images of the target
2
on a very clean and uniform background, suitable for accurate time-resolved photometry (see
Methods). The frames were then visually inspected for artifacts, such as cosmic ray hits and residual
signal from imperfect subtraction of very bright background objects at a location of 1I/'Oumuamua.
This procedure resulted in the rejection of 11 images, a small number compared to the remaining 431
images available for analysis. The restricted dataset was photometrically measured using the
Aperture Photometry Tool21 configured in median sky subtraction mode. Brightness of
1I/'Oumuamua was determined from a circular aperture of a constant 2.26 arcsec (14 pixels)
diameter, which ensured negligible influence of seeing variations and acceptable level of background
noise contribution. We also made similar measurements of brighter background stars, employing this
time a larger aperture of 4.8 arcsec (30 pixels) to account for the elongation of stellar profiles caused
by rapid non-sidereal tracking of the target. Differential photometry of 1I/'Oumuamua was then
reduced to the observing geometry on the first night and absolute calibrated through the background
stars available in the SDSS Photometric catalog22.
In Figure 2 we present a mean combined image of 1I/'Oumuamua, having an effective integration
time of 214.5 min. No signs of cometary activity can be seen, consistent with earlier reports9–12,
however, the sensitivity of our image is a factor of 2.6 to 9.5 higher than those achieved by the
previous searches. Our light curve of 1I/'Oumuamua densely covers a total of 8 hr over two nights.
The median flux corresponds to an observed r'-band magnitude of 22.25 and an absolute magnitude
of 22.02, the latter calculated in the standard way assuming the asteroidal photometric phase
function23 with a parameter G = 0.15 characteristic of solar-system C-types24. Our absolute
magnitude is broadly consistent with the absolute magnitudes reported before10,14. More
interestingly, the range of brightness variations is enormous, at least a factor of ten, or > 2.5 mag,
which exceeds even the largest rotational variations observed in the entire population of 16,419
solar-system asteroids with measured light curves (Figure 1). The light curve of 1I/'Oumuamua was
3
scrutinized for periodicity using the classical Phase Dispersion Minimization algorithm25 with
implemented inverse-variance weighting of the data points26. Periodograms (see example in
Figure 3) systematically show the best data phasing for a frequency of 0.13248±0.00013 hr-1, or
7.5483±0.0073 hr period, implying a double-peaked phased light curve (Figure 4), most easily
explained by shape-dominated brightness variations. Though not as good, a corresponding single-
peaked phasing, consistent with albedo-dominated brightness variations, is also clearly indicated by
the periodograms at approximately twice the double-peaked frequency, 0.26596±0.00026 hr-1, or half
the double-peaked period, 3.7599±0.0037 hr. Other periodicities offer implausible data phasings and
can be safely excluded. We proceed assuming shape-dominated brightness variations and,
consequently, adopt the double-peaked periodicity as 1I/'Oumuamua's synodic rotation period. From
Hapke modeling of the light curve27 we find the long-to-short axis ratio of > 4.63 and the effective
radius of about 80 m (see Methods). The very large elongation together with the measured moderate
rotation rate require a density of > 1034 kg m-3 to prevent the body from falling apart. This limit is
calculated under the assumption that the tensile strength is negligible, which may or may not be true
for 1I/'Oumuamua. Nonetheless, our estimate shows that the body may be strengthless and still have
a density within the ranges of typical Solar System asteroids28, in contrast with the previous
revelations9,10,14.
While the light curve of 1I/'Oumuamua is clearly periodic, it does not repeat exactly from one
rotation cycle to another. As we cannot explain this behavior by instrumental effects, we conclude
that it is a real feature of the light curve, intrinsic to the object. Furthermore, the light curve does not
appear to have a single, unique periodicity because the rotation periods reported by other
studies9,10,13–15 differ from one another and are inconsistent with our data (Figure 5). We recognize
these peculiarities as the characteristic signatures of non-principal-axis (or excited) rotation, also
4
often referred to as tumbling29,30, which has significant consequences for understanding the distant
history of this object.
Although the vast majority of Solar System minor bodies do not show any measurable deviations
from simple rotation, a small fraction of asteroids29,31 and a few comets32,33 have been identified as
tumblers. In particular, comet 1P/Halley was the first minor body found in non-principal-axis
rotation state32 and, subsequently, asteroid (4179) Toutatis was identified as the first tumbler among
asteroids34. Rotational excitation occurs mainly through collisions and, restricted to comets,
sublimation torques. A tumbler then dissipates rotational energy due to stresses and strains resulting
from complex rotation, and returns to simple, minimum-energy rotation state on a certain timescale.
The damping timescale depends on the object's size, shape, density, stiffness, and rotation rate, and
ranges from hundreds years to hundreds of billions years for the known Solar System asteroids35.
Detailed quantification of 1I/'Oumuamua's complex rotation state is beyond the scope of this paper.
Here we explore the fundamental fact that the body was once excited and has not fully relaxed yet.
Evidently, the excitation is not easily explainable by sublimation torques. First, 1I/'Oumuamua does
not show any signs of active outgassing despite a superb sensitivity of our combined image
(Figure 2) and exceptionally favorable orbital circumstances, and second, it is unlikely that the body
was active in the past because the sublimation levels required to excite its rotation would also
generate enormous changes in the rotation rate, quickly leading to rotational instability and
disruption. Instead, the complex rotation state of 1I/'Oumuamua most probably originates from an
impact. Collisional excitation of this body in our Solar System is hardly possible, though, because of
the rarity of collisions even in the (not so) dense main asteroid belt36 and because 1I/'Oumuamua
missed the belt at a safe distance and spent very little time close to the ecliptic plane due to the
5
highly inclined orbit and the large orbital speed. A sensitive non-detection of a body's debris trail
(Figure 2), starkly contrasting with the prominent trail of the similarly-sized impacted asteroid
P/2010 A237, is also not in favor of a very recent collision. Rather, we believe that 1I/'Oumuamua
was excited in another planetary system – presumably its home system – in the distant past. A
damping time scales of typical rubble pile asteroids having the same effective size and rotation rate
as 1I/'Oumuamua, and body's minimum allowable axis ratio, is of about 1 Gyr38. This appears to be
long enough to preserve the signs of collisional excitation over the timescale of a typical interstellar
exile11, supporting our conclusion. Whether 1I/'Oumuamua was excited via a giant collision, giving
birth to this object, or impacted by another small body, is unclear, but both scenarios lead to the same
conclusion that collisional processing of minor body populations in exoplanetary systems is not
uncommon. Supposedly, 1I/'Oumuamua was ejected from its home system with a large number of
similar bodies during a period of dynamical instability1 when frequent collisions are expected.
References
1. Woolum, D. S. & Cassen, P. M. The origin of the Solar System. in Encyclopedia of the Solar System
(eds. Weissman, P. R., McFadden, L.-A. & Johnson, T. V.) 27–28 (1999).
2. Moro-Martín, A., Turner, E. L. & Loeb, A. Will the Large Synoptic Survey Telescope Detect Extra-
Solar Planetesimals Entering the Solar System? Astrophys. J. 704, 733–742 (2009).
3. Sekanina, Z. A probability of encounter with interstellar comets and the likelihood of their existence.
Icarus 27, 123–133 (1976).
4. Stern, S. A. On the number density of interstellar comets as a constraint on the formation rate of
planetary systems. Publ. Astron. Soc. Pac. 102, 793–795 (1990).
5. Jewitt, D. Project Pan-STARRS and the Outer Solar System. Earth Moon Planets 92, 465–476
(2003).
6
6. Francis, P. J. The Demographics of Long-Period Comets. Astrophys. J. 635, 1348–1361 (2005).
7. Cook, N. V., Ragozzine, D., Granvik, M. & Stephens, D. C. Realistic Detectability of Close
Interstellar Comets. Astrophys. J. 825, 51 (2016).
8. Williams, G. MPEC 2017-U181: COMET C/2017 U1 (PANSTARRS). Available at:
https://www.minorplanetcenter.net/mpec/K17/K17UI1.html. (Accessed: 18th November 2017)
9. Meech, K. J. et al. A brief visit from a red and extremely elongated interstellar asteroid. Nature
(2017).
10. Jewitt, D. et al. Interstellar Interloper 1I/2017 U1: Observations from the NOT and WIYN
Telescopes. ArXiv171105687 Astro-Ph (2017).
11. Ye, Q.-Z., Zhang, Q., Kelley, M. S. P. & Brown, P. G. 1I/2017 U1 (`Oumuamua) is Hot: Imaging,
Spectroscopy and Search of Meteor Activity. ArXiv171102320 Astro-Ph (2017).
12. Knight, M. M. et al. The rotation period and shape of the hyperbolic asteroid A/2017 U1 from its
lightcurve. ArXiv171101402 Astro-Ph (2017).
13. Bolin, B. T. et al. APO Time Resolved Color Photometry of Highly-Elongated Interstellar Object
1I/'Oumuamua. ArXiv171104927 Astro-Ph (2017).
14. Bannister, M. T. et al. Col-OSSOS: Colors of the Interstellar Planetesimal 1I/2017 U1 in Context
with the Solar System. ArXiv171106214 Astro-Ph (2017).
15. Feng, F. & Jones, H. R. A. `Oumuamua as a messenger from the Local Association.
ArXiv171108800 Astro-Ph (2017).
16. Marcos, C. de la F. & Marcos, R. de la F. Pole, Pericenter, and Nodes of the Interstellar Minor Body
A/2017 U1. Res. Notes AAS 1, 5 (2017).
17. Mamajek, E. Kinematics of the Interstellar Vagabond A/2017 U1. ArXiv171011364 Astro-Ph (2017).
7
18. Hook, I. M. et al. The Gemini-North Multi-Object Spectrograph: Performance in Imaging, Long-Slit,
and Multi-Object Spectroscopic Modes. Publ. Astron. Soc. Pac. 116, 425–440 (2004).
19. Fukugita, M. et al. The Sloan Digital Sky Survey Photometric System. Astron. J. 111, 1748 (1996).
20. Drahus, M. et al. Fast Rotation and Trailing Fragments of the Active Asteroid P/2012 F5 (Gibbs).
Astrophys. J. Lett. 802, L8 (2015).
21. Laher, R. R. et al. Aperture Photometry Tool. Publ. Astron. Soc. Pac. 124, 737 (2012).
22. Eisenstein, D. J. et al. SDSS-III: Massive Spectroscopic Surveys of the Distant Universe, the Milky
Way, and Extra-Solar Planetary Systems. Astron. J. 142, 72 (2011).
23. Bowell, E., Hapke, B. & Domingue, D. Application of photometric models to asteroids. in Asteroids
II (eds. Binzel, R. & Gehrels, T.) 524 (Univ. Arizona Press, 1989).
24. Luu, J. & Jewitt, D. On the relative numbers of C types and S types among near-earth asteroids.
Astron. J. 98, 1905–1911 (1989).
25. Stellingwerf, R. F. Period determination using phase dispersion minimization. Astrophys. J. 224,
953–960 (1978).
26. Drahus, M. & Waniak, W. Non-constant rotation period of Comet C/2001 K5 (LINEAR). Icarus
185, 544–557 (2006).
27. Hapke, B. Theory of Reflectance and Emittance Spectroscopy. (Cambridge University Press, 2012).
28. Hilton, J. L. Asteroid Masses and Densities. in Asteroids III 103–112 (2002).
29. Pravec, P. et al. Tumbling asteroids. Icarus 173, 108–131 (2005).
30. Samarasinha, N. H. Rotational excitation and damping as probes of interior structures of asteroids
and comets. Meteorit. Planet. Sci. 43, 1063–1073 (2008).
8
31. Warner, B. D., Harris, A. W. & Pravec, P. The asteroid lightcurve database. Icarus 202, 134–146
(2009).
32. Belton, M. J. S., Julian, W. H., Jay Anderson, A. & Mueller, B. E. A. The spin state and
homogeneity of comet Halley's nucleus. Icarus 93, 183–193 (1991).
33. Belton, M. J. S. et al. The complex spin state of 103P/Hartley 2: Kinematics and orientation in space.
Icarus 222, 595–609 (2013).
34. Hudson, R. S. & Ostro, S. J. Shape and Non-Principal Axis Spin State of Asteroid 4179 Toutatis.
Science 270, 84–86 (1995).
35. Pravec, P. et al. The tumbling spin state of (99942) Apophis. Icarus 233, 48–60 (2014).
36. Bottke, W. F. et al. Linking the collisional history of the main asteroid belt to its dynamical
excitation and depletion. Icarus 179, 63–94 (2005).
37. Kim, Y., Ishiguro, M., Michikami, T. & Nakamura, A. M. Anisotropic Ejection from Active
Asteroid P/2010 A2: An Implication of Impact Shattering on an Asteroid. Astron. J. 153, 228 (2017).
38. Breiter, S., Rożek, A. & Vokrouhlický, D. Stress field and spin axis relaxation for inelastic triaxial
ellipsoids. Mon. Not. R. Astron. Soc. 427, 755–769 (2012).
Acknowledgements
M.D., P.G. and B.H. are grateful for support from the National Science Centre of Poland through a
SONATA BIS grant no. 2016/22/E/ST9/00109. We thank the staff of the Gemini Observatory for
assistance and are indebted to the Director of Gemini Observatory for allocating Gemini North time
for this program.
9
Competing Interests
The authors declare that they have no competing financial interests.
Methods
Image cleaning.
In order to obtain accurate time-resolved photometry of 1I/'Oumuamua, it was necessary to carefully
subtract a dense background of stars and galaxies along its trajectory. To remove the background
from a given image, we mean combined up to six background-equalized and flux-normalized images
taken around the same time and subtracted the result of this operation from the reference image. We
did not use the nearest images from a consecutive series to avoid contaminating the photometric
background annulus with the object's signal. An example of the application of this technique to our
1I/'Oumuamua data is presented in Figure S1.
Light curve modeling.
We synthesized the light curve of 1I/'Omuamua with the Hapke approach27, using MIMSA
(modified isotropic multiple scattering approximation) with SHOE (shadow-hiding opposition effect)
and macroscopic roughness influencing reflectance properties. Colorimetric and spectroscopic
observations have revealed that the reflectance spectrum of this object is close to those of the Solar
System C- or D-type asteroids, Trojans, or comets10,11,13,39. We adopted in our model a set of
reasonable literature values for the reflectance parameters of C-type asteroids40, such as a single
scattering albedo of 0.037, an opposition surge amplitude and width of 0.20 and 0.025, respectively,
the asymmetry factor in the Henyey-Greenstein particle phase function equal to -0.47, as well as an
average topographic slope angle of macroscopic roughness of 20°. For simplicity, we assumed that
10
1I/'Oumuamua has a shape of a prolate spheroid. To test how the solution depends on the spin axis
orientation we probed a number of directions with respect to the line of sight (aspect angle θ) and the
Sun-object-observer plane (polar angle φ measured counterclockwise from the object-Sun direction
as seen from observer position). This orientation has been taken for the geometry of sight on 28
October 2017, 00:00 UT and fixed in the external, inertial frame for our whole observing run.
We have checked how the peak-to-peak amplitude of the light curve depends on the axial ratio for
different orientations of the spin pole. A dependency of the flux ratio versus the axial ratio is
presented in Figure S2. Even an enormous amplitude of the light curve does not necessarily mean a
huge axial ratio. We found that to generate a 2.5-mag amplitude, an axial ratio can be as low as 4.63,
inconsistent with the recently published values9,10,13. Large axial ratios were reported with the
assumption that the brightness variation comes purely from cross-section changes. Our lower limit is
a consequence of the assumed reflectance properties of the asteroid surface and partial shadowing
due to over 20° phase angle. When a phase angle is far from 0° and the body is elongated, shadowing
effects as well as asymmetry of the phase function have a strong influence on an observed light
curve. Moreover, for aspect angles smaller than 60° it is virtually impossible to obtain a peak-to-peak
amplitude consistent with 1I/'Oumuamua in spite of blowing up the axial ratio to abnormal values.
Thus we conclude that aspect angle must be larger than 60°.
11
Figures
Figure 1. 1I/'Oumuamua in the context of known minor Solar System bodies. Top: Distribution of the
original reciprocal semi-major axes of the near-parabolic comets41. Middle: Distribution of the light curve
amplitudes of small (left) and all (right) asteroids31. Bottom: Distribution of the rotation periods of small (left)
and all (right) asteroids31. Small asteroids are defined as objects with diameters < 200 m.
12
Figure 2. Deep stack of our r'-band time series, having an effective integration time of 214.5 min. The
negative images of the target to the left and right of the actual positive image were produced by our
background subtraction algorithm and do not affect the photometry. The presented region is 5.5×5.5 arcmin,
North is to the top and East to the left. Despite having a very high surface brightness sensitivity of 28.1 mag
arcsec-2 measured in a 1 arcsec2 region, the image does not show any signs of cometary activity.
13
Figure 3. Phase Dispersion Minimization periodogram calculated with 25 bins and 5 covers for our time
series25,26. Red indicates the best periodicity solution for a frequency of 0.13248±0.00013 hr-1, or
7.5483±0.0073 hr period, consistent with a double-peaked phased light curve.
14
Figure 4. Our data phased for the best periodicity solution, Prot = 7.5483 hr. It is evident that the light curve
does not repeat exactly from one night (27 October 2017 UT) to another (28 October 2017 UT).
15
Figure 5. The same as before but phased for 6.96 hr found by Feng et al.15 (top-left), 7.34 hr found by K.
Meech et al.9 (top-right), 8.10 hr found by B. Bolin et al.13 and M. Bannister et al.14 (bottom-left), and D.
Jewitt et al.10 (bottom-right). None of these four periodicity solutions phase our data, showing that
1I/'Oumuamua does not have a unique rotation period, as expected from tumbling motion.
16
Figure S1. Demonstration of our background cleaning algorithm. Top: example of an original image,
corrected only for overscan, bias, and flatfield. Bottom: clean version of the image above, with stars and
galaxies accurately removed.
17
Figure S2. Flux ratio versus axial ratio for aspect angle θ from 90° to 30°. For each angle two different
trajectories are presented, one for minimum and one for maximum flux ratios over a series of polar angles φ
for a given aspect angle θ.
Supplementary References
39. Masiero, J. Palomar Optical Spectrum of Hyperbolic Near-Earth Object A/2017 U1. (2017).
40. Helfenstein, P. & Veverka, J. Physical characterization of asteroid surfaces from photometric
analysis. in 557–593 (1989).
41. Królikowska, M. Warsaw Catalogue of cometary orbits: 119 near-parabolic comets. Astron.
Astrophys. 567, A126 (2014).
18
|
1612.08635 | 2 | 1612 | 2017-01-07T08:27:03 | Collisional parameters of planetesimal belts, precursor of debris disks, perturbed by a nearby giant planet | [
"astro-ph.EP"
] | Planetesimal belts are invoked to explain the prolonged existence of debris disks. Important parameters to model their collisional evolution and to compute the dust production rate are the intrinsic probability of collision $P_i$ and the mean impact velocity $U_c$. If a planet orbits close to the belt, the values of both these parameters are affected by its secular perturbations yielding a strong correlation between eccentricity $e$ and pericentre longitude $\varpi$. We adopt a new algorithm to compute both $P_i$ and $U_c$ in presence of various levels of secular correlation due to different ratios between proper and forced eccentricity. We tested this algorithm in a standard case with a Jupiter--sized planet orbiting inside a putative planetesimal belt finding that it is less collisionally active compared to a self--stirred belt because of the $e - \varpi$ coupling. The eccentricity of the planet is an important parameter in determining the amount of dust production since the erosion rate is 10 times faster when the planet eccentricity increases from 0.1 to 0.6. Also the initial conditions of the belt (either warm or cold) and its average inclination strongly affects $P_i$ and $U_c$ and then its long term collisional evolution in presence of the planet. We finally apply our method to the planetesimal belts supposedly refilling the dust disks around HD 38529 and $\epsilon$ Eridani. In the most collisionally active configurations, only a small fraction of bodies smaller than 100 km are expected to be fragmented over a time--span of 4 Gyr. | astro-ph.EP | astro-ph |
Mon. Not. R. Astron. Soc. 000, 1 -- 17 (.....)
Printed 14 October 2018
(MN LATEX style file v2.2)
Collisional parameters of planetesimal belts, precursor of
debris disks, perturbed by a nearby giant planet.
F. Marzari1 and A. Dell'Oro2
1Dept. of Physics, University of Padova, 35131 Italy
2INAF, Osservatorio Astrofisico di Arcetri, Largo E. Fermi 5, I-50125, Firenze, Italy
Accepted .....; Received ..... ; in original form ........
ABSTRACT
Planetesimal belts are invoked to explain the prolonged existence of debris disks.
Important parameters to model their collisional evolution and to compute the dust
production rate are the intrinsic probability of collision Pi and the mean impact ve-
locity Uc. If a planet orbits close to the belt, the values of both these parameters are
affected by its secular perturbations yielding a strong correlation between eccentricity
e and pericentre longitude . We adopt a new algorithm to compute both Pi and
Uc in presence of various levels of secular correlation due to different ratios between
proper and forced eccentricity. We tested this algorithm in a standard case with a
Jupiter -- sized planet orbiting inside a putative planetesimal belt finding that it is less
collisionally active compared to a self -- stirred belt because of the e − coupling. The
eccentricity of the planet is an important parameter in determining the amount of
dust production since the erosion rate is 10 times faster when the planet eccentricity
increases from 0.1 to 0.6. Also the initial conditions of the belt (either warm or cold)
and its average inclination strongly affects Pi and Uc and then its long term collisional
evolution in presence of the planet. We finally apply our method to the planetesimal
belts supposedly refilling the dust disks around HD 38529 and ǫ Eridani. In the most
collisionally active configurations, only a small fraction of bodies smaller than 100 km
are expected to be fragmented over a time -- span of 4 Gyr.
Key words: planetary systems; planets and satellites: dynamical evolution and sta-
bility
1
INTRODUCTION
Debris disks, mostly detected in thermal infrared, are ob-
served around 10 -- 20% of solar type stars (Hillenbrand et al.
2008; Trilling et al. 2008; Sibthorpe et al. 2013). The
micron -- sized grains populating these disks are short lived
mostly because of radiation related forces and collisional ero-
sion. As a consequence, their refilling requires the presence
of a reservoir of planetesimals whose collisions constantly
produce new dusty debris. Indeed debris disks are the only
signature of these belts whose properties like size distribu-
tion, mechanical strength and dynamical excitation cannot
be constrained by observations. The only way to link plan-
etesimals and dust is through numerical models reproducing
the SED and resolved images of debris disks. However, these
models not always give unique solutions.
Analogues of the solar system asteroid or Kuiper belt,
planetesimal belts are leftovers from the planet forma-
tion process and they eventually continue to evolve un-
der mutual collisions. The frequency of debris disks as
well as the infrared excess strength decline with the stel-
lar age (Krivov 2010) suggesting that the collisional pro-
cess slowly grinds down the initial planetesimal population.
Different mechanisms have been invoked to stir up a plan-
etesimal swarm igniting mutual destructive collisions like
self -- stirring by larger planetesimals (Chambers & Wetherill
2001; Kenyon & Bromley 2004) and perturbations by plan-
ets (Mustill & Wyatt 2009). Concerning this last mecha-
nism, it is still controversial from an observational point
of view if there is any significant correlation between stars
with dust emission and the presence of known planets
(Bryden et al. 2009; Moro-Mart´ın et al. 2007a) even if some
of such systems are already known like that in ǫ Eridani
(Benedict et al. 2006a), Fomalhaut (Kalas et al. 2008) and
HD 38529 (Moro-Mart´ın et al. 2007b). From a theoretical
point of view, it is expected that debris disks and plan-
ets coexist in a large number of systems being both the
outcome of dust coagulation and planetesimal accumulation
even if debris disks appear to be more common than mas-
sive planets (Moro-Mart´ın et al. 2007a). This may be due to
2
the presence of debris disks even in systems where the plan-
etesimals were not able to form the core of giant planets. In
this scenario, the dust would be produced in mutual colli-
sions possibly triggered by self -- stirring. There are significant
observational problems in statistically assessing the proba-
bility of finding giant planets and debris disks coexisting in
the same system. Young stars have a higher percentage of
debris disks but radial velocity surveys for planets around
them are difficult because of the noise due to stellar activity.
On the other hand, planets can be easily detected around
old stars where debris disks may have been eroded away by
collisional evolution and are expected to be less common. It
is also difficult to precisely locate the debris belts in radial
distance in order to asses the potential detectability of close-
by planets. Finally, it should be noted that the inventory of
giant exoplanets is still poor for semimajor axes larger than
a few AU.
Assuming that in some systems different dynamical
mechanisms lead to a configuration with a massive planet
in an inside orbit respect to a debris disk, this would be
a particularly interesting architecture since the disk struc-
ture is very sensitive to the planet gravitational perturba-
tions potentially driving the formation of arcs, gaps, warps
and asymmetric clumps in the disk (Moro-Mart´ın et al.
2005, 2007b). Even the gap between two components debris
disks is suspected to be carved by intervening planets scat-
tering away the remnant planetesimals (Su & Rieke 2014;
Shannon et al. 2016) and a survey based on direct imag-
ing is searching for planets in systems with double debris
disks (Meshkat et al. 2015). In addition to these direct ef-
fects, a planet also perturbs the planetesimal belt stirring
up their orbits and affecting the collision probability and
impact velocity (Mustill & Wyatt 2009). The level of stir-
ring depends on the planet mass, vicinity to the belt, either
inside or outside, and orbital eccentricity. An excited plan-
etesimal belt may give origin to a brighter debris disk at
start (Wyatt et al. 2007), but its lifetime will be significantly
shorter due to the collisional erosion.
We can envisage different scenarios related to the co-
existence of planets and debris disks. A single giant planet,
which had a limited amount of inward migration, can clear
part of the leftover planetesimal population leaving an inner
cavity in the planetesimal disk. In this case we would find
a low eccentricity planet secularly perturbing an external
planetesimal belt. This configuration does not exclude the
presence of additional planets in the system moving in inside
orbits. More complex dynamical configurations are expected
when a multi -- planet system undergoes a period of extended
chaotic evolution characterised by planet -- planet scattering.
In extreme cases this dynamical evolution may lead to a
complete clearing of local planetesimal belts because of the
intense perturbations of the planets on highly eccentric or-
bits (Bonsor et al. 2013; Marzari 2014). However, less vio-
lent evolutions may lead to a final architecture where a giant
planet on an eccentric orbit is located inside, or outside, the
surviving belt and it affects its evolution via secular pertur-
bations. The two different mechanisms may lead to a wide
range of possible eccentricities for the planet perturbing the
belt.
When a planetesimal belt is accompanied by a close-
by planet, the relative impact velocity and frequency of
collisions of the perturbed belt, used to predict its long
term collisional evolution, cannot be calculated with the
Opik/Wetherill analytic formulation (Opik 1951; Wetherill
1967) even in the improved formulations developed to study
the evolution of the asteroid belt (see Davis et al. (2002)
for a review). In presence of a massive planet, in particular
if on a highly eccentric orbit, the forced component in the
secular evolution of the planetesimal orbits may lead to a
strong correlation between eccentricity and perihelion longi-
tude. This correlation invalidates the Opik/Wetherill meth-
ods based on a uniform distribution of the orbital angles
(perihelion longitude and node longitude) derived under the
assumption of periodic circulation of the two angles. These
methods work properly when describing the evolution of the
asteroid or Kuiper belt in the solar system where the forced
eccentricity is significantly smaller than the proper one, but
in dynamical configurations where the planetesimal ring is
perturbed by a giant planet on a highly eccentric orbit they
fail.
A first attempt to overcome this problem is described in
Mustill & Wyatt (2009) where they first derive refined val-
ues of the forced eccentricity from which they compute an
average value of the impact velocity between the planetes-
imals multiplying the forced eccentricity by the local Kep-
lerian velocity and a constant value c ∼ 1.4. In this paper
we apply an innovative semi-analytic method to estimate
both the intrinsic probability of collision Pi i.e. the collision
rate per unit cross -- section area of target and projectile per
unit time, and the average collision speed Uc, derived from
a detailed statistical frequency distribution of the relative
velocities, in a planetesimal belt perturbed by an eccentric
giant planet. This method is designed in order to fully ac-
count for the different levels of correlation between eccentric-
ity and perihelion longitude of the planetesimals caused by
the secular dynamics. Both Pi and Uc are needed to prop-
erly model the collisional evolution of a planetesimal disk
and predict the dust production rate refilling the related de-
bris disk. In particular, Pi is more relevant in establishing
the amount of erosion of the belt. We consider two distinct
plausible scenarios where, prior to the evolution of the per-
turbing planet close to the belt, the planetesimal disk was
either dynamically non -- excited (cold population) or excited
(warm population). In the former case (cold population),
when the planet approaches the belt during its evolution,
the proper eccentricity of the planetesimals is immediately
excited to a value approximately equal to the secular forced
one (Th´ebault et al. 2006). This is also the scenario explored
by Mustill & Wyatt (2009) and it is the dynamical configu-
ration where the correlation between e and is maximised.
If instead the eccentricities of the planetesimal population
were already significantly excited when the planet sets in and
begins to perturb the belt (warm belt) a different dynam-
ical configuration is achieved where the proper and forced
eccentricities may significantly differ. We will consider cases
where the eccentricity, prior to the onset of the secular per-
turbations, is an increasing fraction of the forced eccentric-
ity. In these cases after the onset of the secular perturbations
the degree of correlation between e and will be less robust
within the belt and pseudo -- librator states will appear in the
population influencing the values of both Pi and Uc. While
modelling a warm belt, it is reasonable to expect that also
the inclination is excited as well so we will explore the ef-
fect of a high planetesimal inclination on both the collisional
parameters.
If a belt is densely populated by planetesimals, a sig-
nificant collisional damping of the eccentricity may occur
(Stewart & Wetherill 1988) leading it to a state of cold belt
before the planet approaches. After the planet sets into a
perturbing orbit, the collisional damping may still be effi-
cient in reducing the eccentricity, but in this case it will
affect only the proper term causing its progressive decrease.
We explore also this scenario and estimate the values of the
collisional parameters to compare with non -- damped cold
and warm belts.
In Section 2 we briefly summarise the expected dynam-
ical behaviour of a planetesimal belt perturbed by a planet.
In Section 3 we outline the method developed to compute
the average values of intrinsic probability of collision Pi and
impact velocity Uc for a belt characterised by the above men-
tioned secular dynamics. In Section 4 we apply the method
to a 'standard' case to illustrate the effects of the secular dy-
namics and compare the predicted values of Uc with those
derived by Mustill & Wyatt (2009). We also derive and com-
pare the values of the collisional parameters in cold and
warm belts with and without inclination excitation. In Sec-
tion 5 we model the collisional evolution of a putative belt
with the previously estimated collisional parameters while
in Section 6 we model a real system, HD 38529, and com-
pare its evolution with that of ǫ Eridani. Finally, in Section
7 we summarise and discuss our results.
2 SECULAR EVOLUTION AND
CORRELATION BETWEEN
ECCENTRICITY AND PERIHELION
LONGITUDE
The dynamical evolution of a minor body population per-
turbed by a planet is classically described by the linear secu-
lar theory of Laplace -- Lagrange (Murray & Dermott 1999).
For a mass -- less planetesimal population, the evolution with
time of the non -- singular variables h = e sin ω and k =
e cos ω, where e and ω are the eccentricity and longitude of
the pericentre of the osculating orbit, respectively, is given
by
h = epsin(At + B) + ef
k = epcos(At + B)
(1)
where B is a constant determined by the initial condi-
tions of the system, ep is termed proper eccentricity while
ef is the forced one. In our case we consider the simpli-
fied situation in which the planet is not perturbed by ad-
ditional bodies so that ef is constant. In Eq. 1 we assume
that the reference frame for the computation of the orbital
elements is aligned with the apsidal line of the planet or-
bit. Approximate values for the proper frequency A and the
forced eccentricity ef can be derived from the simplified lin-
ear disturbing function (Murray & Dermott 1999) and are
given in Mustill & Wyatt (2009)
Collisional parameters of planetesimal belts
3
ef ∼
5
4
A ∼ n
αepl
3
4
mpl
ms
α2 ¯α
(2)
where ms is the mass of the central body (the star),
epl the eccentricity of the planet and n is the mean motion
of the test body. For a configuration in which the planet is
interior to the planetesimal orbit α = apl
a and ¯α = 1 while
α = a
and ¯α = α for an exterior planet. a is the semimajor
apl
axis of the planetesimal and apl the semimajor axis of the
planet (Murray & Dermott 1999; Mustill & Wyatt 2009).
This classical perturbation theory works far from mean mo-
tion and secular resonances and, as a consequence, some
values of α lead to incorrect predictions. In addition, the
theory has been developed to second order in the eccentric-
ity and inclination of the bodies and it is a good approx-
imation only for small values of these orbital parameters.
Among known extrasolar planets, a significant fraction have
high orbital eccentricities for which either semi -- numerical
approaches (Michtchenko & Malhotra 2004) or higher or-
der theories (Libert & Henrard 2005) are preferable. How-
ever, a numerical exploration of the reliability of the for-
mulas of the linear theory even for large values of epl by
Mustill & Wyatt (2009) has shown that the precession rate
can be well described by the disturbing function developed
by Heppenheimer (1978) for binary stars:
A ∼ 2πn
3
4
mpl
ms
α2 ¯α
4(1 − e2
pl)3/2
(3)
for values of epl beyond 0.2 even if the correction proposed
by Th´ebault et al. (2006) was not tested. This last is a more
accurate prescription for the frequency of the secular os-
cillations induced by a binary companion empirically de-
rived from several numerical simulations. In our context, a
very precise value of the frequency A is not strictly required
since what matters is that the randomisation of the angles
is achieved. A constant frequency of circulation is the only
assumption about θ = At + B used in our statistical model.
The forced eccentricity ef is instead well approximated by
the linear secular theory even for large values of epl.
To test the reliability of the linear secular theory in
dealing with the main dynamical features of a planetesimal
belt perturbed by an eccentric planet, we have compared the
outcome of a short term numerical simulation with the pre-
dictions of the theory. In Fig.1 we show the evolution with
time of a population of planetesimals orbiting between 15
and 20 AU on initially circular orbits and perturbed by a
planet at 5 AU with an eccentricity epl = 0.5. The 15th order
RADAU integrator (Everhart 1985) has been used for this
short term simulation. In the upper panel we show the h and
k variables after about 10 Myr of evolution. The pericentre
longitude is randomised since the period of the secular cir-
culation is lower than 1 Myr and the integration time -- span
is also significantly longer than the 'crossing time' defined
in Mustill & Wyatt (2009) for an internal perturber. This
time is given by:
4
tcross ∼ 1.53 × 103 (1 − e2
epl
M⊙(cid:19)−1
×(cid:18) mpl
(cid:18) M∗
M⊙(cid:19)
1
2
yr,
(4)
pl)3/2
2 ×
9
(cid:16) a
10AU (cid:17)
×(cid:16) apl
1AU(cid:17)−3
and, for our test belt, the longest tcross is about 3.6 × 105
yr.
The theoretical evolution for e and predicted by the
linear secular theory can be derived combining the two vec-
tors drawn in the plot representing the forced ef and proper
ep eccentricities. In the bottom panel the planetesimal ec-
centricity is shown as a function of the perihelion longi-
tude and compared to the theoretical curve. The ana-
lytic curves are computed for a single value of semimajor
axis and, as a consequence, they only match the evolution
of the planetesimals with initial semimajor axis similar to
that used in the computation of the curves. The agreement
between the linear theory and the numerical results is really
good apart from a few scattered points which are due to the
effects of mean motion resonances with the planet and can
be neglected.
In the figures, we have assumed that ep ∼ ef , condi-
tion which is based on the premise that, before being per-
turbed by the planet, the planetesimals were initially on
unperturbed almost circular orbits with ep0 ∼ 0. The ap-
proximation ep ∼ ef is common when exploring the pertur-
bations of a massive body on an initially cold planetesimal
belt (Heppenheimer 1978; Whitmire et al. 1998). In partic-
ular, if the planet is injected on a highly eccentric orbit after
a period of dynamical instability, the belt of planetesimals
will suddenly feel the strong secular perturbations of the
planet. If we assume that as 'time 0' of the secular pertur-
bations the belt was cold with ep0 ∼ 0, then the condition
ep ∼ ef is naturally imposed in the subsequent evolution.
We term here ep0 the proper eccentricity before the onset
of the planet perturbations. In this case it would not be
correct to call it a 'proper' eccentricity since it is just the
average eccentricity of the cold belt. However, since it will
determine the subsequent value of the proper eccentricity
once the planet begins to perturb the belt, hereinafter we
will use ep0 to indicate it.
A scenario different from that of an initial cold belt
is produced by a belt which has been excited before the
arrival of the perturbing planet, a warm belt. Various
mechanisms may contribute to the dynamical excitation
like stellar flybys, self -- stirring and planet stirring related
to an extended period of planet -- planet scattering (see
Matthews et al. (2014) for a review). This last mechanism,
due to its large variety of possible outcomes, may activate
the planetesimal belt inducing high eccentricities and incli-
nations before a planet is deposited on a finally stable orbit
close to the belt (Bonsor et al. 2013; Marzari 2014). In a
warm belt the initial planetesimal eccentricity ep0 is signifi-
cantly higher than 0 at the time of the planet approach. The
secular modelling of an initially warm belt is more complex
respect to that of a cold belt since in the former case the
initial conditions for the secular evolution are different from
the simple assumption ep ∼ ef . In Fig.2 we show how the
new proper eccentricity ep of each planetesimal in the belt,
after the planet perturbations are switched on, can be com-
h
y
t
i
c
i
r
t
n
e
c
c
E
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
ep
θ
ef
-0.2 -0.15 -0.1 -0.05 0 0.05 0.1 0.15 0.2
k
0.4
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
0
50
100 150 200 250 300 350
Per. Long. (deg)
Figure 1. Secular evolution of a population of planetesimals mov-
ing in a ring with semimajor axis ranging from 15 to 20 AU
and initially circular orbits. They are perturbed by a Jupiter --
size planet with apl = 5 AU and eccentricity epl = 0.5. In the
top panel the evolution of the non -- singular h and k variables is
illustrated (h = e cos(), k = e sin()). The arrows show the
forced ef and proper ep eccentricities for a = 15 AU (in blue) and
a = 20 AU (in black), which, once combined, give the eccentricity
of the planetesimals. The blue and black dashed lines show the
curves obtained by combining proper and forced eccentricity. On
the bottom panel the correlation between and e is highlighted.
puted from ep0 (initial average eccentricity of the belt before
the approach of the planet) and how the secular theory can
be used to predict the subsequent dynamical evolution. In
the upper panel of Fig.2 we sketch the mode of operation of
the secular theory using vector formalism. The blue circle
marks the location of the tips of all the initial proper ec-
centricity vectors ep0's of the warm belt. All these vectors
have the same modulus (we assume that all planetesimals
have the same initial ep0 equal to the average of the belt)
while their orientation is random depending on . When
the planet begins to perturb the belt, a common forced ec-
centricity vector ef appears in the secular evolution. As a
consequence, each body will acquire a new proper eccen-
tricity vector ep which depends on the the initial proper
eccentricity vector ep0 through the relation ef + ep = ep0.
The new proper eccentricity vector ep = ep0 − ef will mark
the secular evolution of each planetesimal from then on. The
modulus of the new ep ranges from ef − ep0 and ef + ep0
with all intermediate values being possible, if the belt is
h
e
ep
e f
e
ep0
ep
ep0
k
0.6
0.4
h
0.2
0
-0.2
-0.4
-0.2
0
k
0.2
0.4
0.5
0.4
0.3
0.2
0.1
e
0
0
60
120
180
240
300
360
w (deg)
Figure 2. In the upper panel the secular dynamics is sketched.
The blue circle describes the location of the tip of the proper
eccentricity vector ep0 in an unperturbed warm belt. When the
planet perturbs the belt, new initial proper vectors ep are com-
puted joining the tip of ef with the tip of ep0. Once the initial
eps are computed in this way, the secular evolution can be pre-
dicted following the circle centred on the tip of ef and depicted
by the tip of ep. In the middle and lower panels the secular be-
haviour is compared to the outcome of numerical integrations.
We select 5 orbits with semimajor axis equal to 15 AU, initial
values ep0 = 0.5 · ef and 5 different values of the initial pericentre
longitude 0o, 45o, 90o, 135o, and 180o. The analytic predictions
are given by the dashed continuous lines and they match closely
the numerical results drawn by different colours depending on the
initial value of . In both panels, for small values of ep, pseudo --
librations around 0o are observed (red and green points).
Collisional parameters of planetesimal belts
5
enough crowded. The ep's will be used, together with ef , to
compute the subsequent secular evolution featured by circles
centred on the tip of ef and with radius equal to ep. In the
middle and lower panels of Fig.2 we compare the analytic
predictions of the secular theory with numerical integration
of five different planetesimal orbits. They are all started with
semimajor axis a = 15 AU and all have the same value of
initial eccentricity ep0 = 0.5 · ef . The pericentre longitude
is set to 0o, 45o, 90o, 135o, and 180o, respectively. In this
way we cover the most relevant secular behaviours typical
of an initially warm belt perturbed by a planet. For small
values of ep a pseudo -- libration around 0o is observed while
for larger values circulation is restored. In all cases there is
a significant correlation between eccentricity and pericentre
longitude.
In modelling warm belts we will sample three different
initial ratios between ep0 and ef i.e. 0.25, 0.5, and 1. Link-
ing ep0 to ef is an arbitrary choice but it is adopted to give
an idea of how higher initial values of ep0 influence both Pi
and Uc and it appears more robust than selecting random
values of ep0. It is less arbitrary, it can be easily replicated
and, in addition, it includes a radial dependence which may
be present in the initial belt. At present we do not have
the means to reconstruct the past history of a warm belt
and of the dynamical mechanisms which may have stirred it
(planet formation, protoplanets roaming around, a phase of
planet-planet scattering etc...) and, as a consequence, many
different initial distributions of the planetesimal proper ec-
centricities are conceivable. We focus in this paper on those
where the proper eccentricity does not exceed the forced one
and decreases with radial distance. For warm belts we will
explore a scenario where the dynamical excitation involves
also the inclination modelling, together with the case with
i = 3o, one with i = 15o.
If the planetesimal belt is densely populated, a strong
collisional damping may occur even after the planet reached
its close-by orbit. This damping is due to the loss of orbital
energy after each collision and, on average, it causes a reduc-
tion of the proper eccentricity ep. Any crowded belt might
be drag to a condition similar to that illustrated in Fig. 3
even in presence of a perturbing planet. The proper eccen-
tricity is smaller than the forced one and all planetesimals
evolve in pseudo -- librating orbits with a degree of alignment
which depends on the ratio between proper and forced ec-
centricity. We will model also this dynamical configuration
and compute values of Pi and Uc in this scenario.
We have to point out that our estimates of Pi and Uc
cannot be applied to very young belts, since our model re-
quires a full randomisation of the pericentre of the planetesi-
mals. This occurs on a timescale of some tens of Myrs when
the perturber is a Jovian -- type planet limiting the appli-
cability of our computations to older disks. From Eq. (3)
we can deduce that smaller size planets need more time
to randomise the pericentre since the secular period is in-
versely proportional to the planet mass mpl. For example, a
Neptune -- size planet would lead to pericentre randomisation
on a timescale approximately 20 times longer i.e. about half
of a Gyr. Our estimates of Pi and Uc will be valid only after
that time.
6
Figure 3. If the planetesimal belt is densely populated, a signif-
icant collisional damping may occur reducing the proper eccen-
tricity even after the planet ignited its secular perturbations. In
this case the secular evolution will lead to pseudo -- librations of
the pericentre of all bodies and the secular evolution is described
by the dashed circle.
3 THE METHOD
An essential prerequisite to model the collisional evolution
of a population of minor bodies, like planetesimals, is a
quantitative estimate of two parameters: the average im-
pact velocity Uc and the intrinsic probability of collision
Pi. The impact velocity appears in all basic steps determin-
ing the outcome of a collision and it discriminates between
accumulation, cratering or shattering events. Together with
semi-empirical scaling laws describing the strength of a body
suffering a collision, Uc determines the mass of the largest
remnant body and the size distribution of the escaping frag-
ments. The intrinsic probability of collision Pi is related to
the frequency of collisions between the bodies of the given
population. It is a property related to the orbital distribution
and defined independently from the number or size distribu-
tion of the bodies populating the belt. For example, for two
given orbits, Pi is defined as the mean number per year of
close encounters with minimum distance less than 1 km oc-
curring between two points moving along the two orbits. If
two bodies with radius RT ("target") and Rp ("projectile")
move along the above mentioned orbits, the mean number of
collisions per unit of time is Pi(RT + Rp)2. In wider terms,
if the orbit of a target body with radius RT moves within a
swarm of Np projectiles with radius Rp, and Pi,k is the in-
trinsic probability of collision between the target orbit and
the orbit of the k-th projectile, the mean number of collisions
per unit of time suffered by the target is Pk Pi,k(RT +Rp)2,
or:
dn
dt
= hPii(RT + Rp)2Np
(5)
where hPii = (Pk Pi,k)/Np is the mean intrinsic probability
of collision (Farinella & Davis 1992; Davis et al. 2002). In
this paper we will treat only cases of targets impacted by
populations of projectiles, so hereinafter we use the symbol
Pi always with the meaning of mean intrinsic probability of
collision. It has been computed for different populations of
minor bodies in the solar system and, for instance, its mean
value for the asteroid belts is Pi ∼ 2.9 × 10−18 km−2 yr−1
(Bottke et al. 2002). The important aspect of Pi is that it is
not simply a particle -- in -- a -- box model where the bodies move
freely between collisions, but it accounts for the Keplerian
dynamics of the bodies even when they are on perturbed
orbits.
In the case of planetesimal belts perturbed by a planet
in a close-by orbit, the classical Opik/Wetherill approach for
the computation of Pi and Uc cannot be used if the forced
eccentricity is not significantly smaller than the proper ec-
centricity. The Opik/Wetherill statistics, hereinafter termed
"canonical" statistics, is based on the following assumptions:
(1) the semimajor axes a, eccentricities e and inclinations
I of the osculating orbits are fixed;
(2) the rate of variation of the node longitudes Ω and peri-
centre arguments ω of the osculating orbits are constant. In
other words, the osculating orbits circulate uniformly;
(3) the motion of the planetesimals along their osculating
orbits is fully described by the second Kepler's law;
1998); Dell'Oro et al.
The first attempt to overcome the assumptions of the
canonical statistics has been done by Dell'Oro & Paolicchi
(1997,
particular
(1998) developed a mathematical
Dell'Oro & Paolicchi
formalism for the study of the statistics of collisions
among asteroids provided that the following more general
assumptions are fulfilled:
(1998).
In
(1) the orbital parameters a, e and I of the osculating orbits
are fixed
(4) none of the osculating elements a, e and I is correlated
with one of the angular elements Ω, ω or mean anomaly M
Indeed, thanks to its numerical implementation, the
method of Dell'Oro & Paolicchi (1998) can be used even
if condition (1) is not fulfilled by simply substituting
the ensemble of fixed values of a, e and I with a ficti-
tious list of "child" elements a, e and I whose distribu-
tions fit ad-hoc parent distributions. So, in general, in all
cases where condition (4) is fulfilled, the method proposed
by Dell'Oro & Paolicchi (1998) can be employed. Unfortu-
nately, the dynamical behaviour of a planetesimal belt per-
turbed by a planet strongly violates condition (4) because
of the strong correlation between eccentricity and perihelion
longitude, entailing a correlation among e, Ω and ω. To ac-
count for this correlation, we envisioned a different technique
described here below.
Our new algorithm adopts a Monte Carlo approach con-
sisting in a random exploration of the phase space of posi-
tions and velocities of the bodies in order to derive their
rate of collision and impact velocity distribution. It can be
summarised in four consecutive steps.
In the first step, a numerical model of the disk is pre-
pared by generating a list of bodies with semimajor axis
randomly and uniformly chosen in an interval [amin, amax].
Different radial distributions can be implemented, but in
this paper we adopt this simple assumption which implies
a decrease of the planetesimal superficial density as r−1.
The forced eccentricity for each body is computed from Eq.
2 (Mustill & Wyatt 2009) while the inclination I is ran-
domly generated from a uniform distribution in the inter-
val [0, Imax]. Finally, a value of proper eccentricity ep is
computed for all bodies which is derived from the initial
assumed value of proper eccentricity ep0 of each planetes-
imal in the belt before it is perturbed. In all our models,
the average number of bodies (orbits) used in the Monte
Carlo statistics is of the order of 10, 000. Our final goal is to
compute the statistical parameters Pi and Uc for collisions
occurring between all bodies belonging to our list, consid-
ered representative of the structure of the disk, and some
selected target bodies (tracers) whose orbits have inclina-
tion IT = Imax/2, semimajor axes aT chosen at fixed steps
between [amin, amax], and values of ep, ef computed with
the same rules for the disk bodies. In this way, Pi and Uc
are functions of the semimajor axis of the target only and
can be evaluated at different locations on the disk. In the
case of a warm belt, since the tracers may have different
values of ep, 1000 tracers are used for each semimajor axis
and the values of Pi and Uc are computed as average over
all of them.
The second step consists of a random sampling of the
mean anomaly M and of the secular angle θ, the angle be-
tween the vectors describing the proper and forced eccentric-
ity, respectively (see Fig. 1). Both samplings are performed
assuming a uniform distribution of the angles since we know
from the secular theory that θ precesses at a regular pace,
at least as a first approximation, while the mean anomaly
M circulates with the constant frequency n. From the val-
ues of ef , ep and θ we can derive the osculating eccentricity
completing the set of orbital elements. From the long list
of orbital elements sampled for each body of the disk and
tracers we can compute the list of positions and velocities
(rt, vt, rp, vp) of projectiles and targets. This two-steps sam-
pling allows to reproduce in the phase space, described by
positions and velocities, the distribution of the orbital ele-
ments of projectiles and targets imposed by the secular per-
turbations of the planet. This distribution is characterised
by different levels of correlation between e and ω which de-
pend on the initial choice of ep0. At this point we have a
model disk reflecting the secular dynamics in Cartesian co-
ordinates.
The third step consists in computing, from the distri-
bution of the positions and velocities of both projectiles and
targets, the rate of close approaches and the distribution of
impact velocities. For each target the number ν(R) of close
encounters within a given distance R per unit time is, by
definition, the ratio:
ν(R) =
N (R, T )
T
(6)
where N (R, T ) is the number of close encounters occurred
during an interval of time T and within a distance R. It is
noteworthy that, in the context of the above equation, R is
the distance between the centres of the two bodies during the
approach and it is not the radius of either of the two bodies.
Assuming a general dynamical stability of the system and
choosing T long enough, the number ν(R) no longer depends
either on the initial conditions of the system or on the value
of T . We can rewrite this ratio as:
ν(R) =
N (R, T )
S(R, T )
S(R, T )
T
(7)
Collisional parameters of planetesimal belts
7
where S(R, T ) is the sum of the durations of all close en-
counters occurred during the interval of time T and within
the distance R. The ratio:
τ (R) =
S(R, T )
N (R, T )
(8)
is the mean value of the durations of the close encounters
within a distance R, while:
p(R) =
S(R, T )
T
(9)
is the probability to find a projectile within a distance R
from the target at a randomly chosen instant of time. The
mean close encounter duration τ (R) is related to the prop-
erties of the relative motion between projectile and target.
Assuming that the relative trajectory of the bodies can be
approximated as a rectilinear motion with constant veloc-
ity v = vt − vp, the average (expected) duration of a close
encounter is (4/3)(R/v), taking into account that the projec-
tile can pass anywhere inside the sphere of radius R around
the target. This rectilinear motion approximation is valid
only for values of the close approach distance R significantly
smaller than the size of the orbits of the planetesimals. The
probability p(R) is directly derived from the (rt, vt, rp, vp)
list by computing the ratio between the number of samples
for which rt − rp < R and the total number of samples. In
this way the rate of the close encounters can be expressed
in terms of the sampled quantities as:
ν(R) =
3
4
1
N R Xk
vk
(10)
where vk is the relative velocity of the k-th sample of target-
projectile pairs. The sum includes only those cases for which
rt − rp < R while N is the total number of samples
(Dell'Oro 2016).
The rate ν(R) is derived as a function of the close ap-
proach distance R. By definition of intrinsic probability of
collision, Pi = ν(R) if R = 1 km. On the other hand, due to
the numerical limitations of the Monte Carlo approach, it is
not possible to evaluate the rate ν(R) directly for R = 1 km,
since it requires an excessive computational effort. Neverthe-
less, for values of R small compared to the linear dimensions
of the orbits, the rate ν is expected to be proportional to
the geometric cross section, that is ν(R) ∝ R2 (the effect of
the gravitational focusing is negligible due to the small sizes
of the planetesimals). For all the cases investigated in this
work we have verified that ν(R) is really proportional to R2
when R is small enough. For this reason, we extrapolate the
function ν(R) down to R = 1 km by assuming that the R2
trend is maintained. In short, within the interval of values
of R for which ν(R) results to be proportional to R2, the
ratio ν(R)/R2 provides the value of Pi directly.
Together with Pi, we also compute the mean v(R) of the
relative velocity for all close encounters with rt − rp < R.
The parameter v(R) is correctly evaluated on the basis of
the (rt, vt, rp, vp) list as:
where again vk is the relative velocity of the k-th target-
(11)
k
v(R) = Pk v2
Pk vk
8
projectile pair and the sum includes only those cases for
which rt − rp < R. From Eq. 10 it can be deduced that
each of the target-projectile pairs contributes to the final
evaluation of the frequency ν(R) with a term vk/R. This
comes from the fact that each orbital geometry has its own
probability to occur, as outlined by Bottke et al. (1994). In
other words, each value of the relative velocity vk of the
planetesimal ensemble has to be weighted by vk itself in
order to obtain the correct probability distribution of the
relative velocities. The same holds true for any other pa-
rameter related to close encounters. But unlike ν(R), the
value of v(R) is expected to tend asymptotically to a finite
value for R smaller and smaller. This limit is the mean value
Uc of the impact velocity. In general, for the cases under in-
vestigation in this work, we have verified that the function
v(R) decreases for lower R and it becomes constant from a
certain value onward, providing our estimation of Uc.
The procedure described above leads to an accurate es-
timate of both Pi and Uc (Dell'Oro 2016). The algorithm has
been extensively validated on different sets of test cases and,
in particular, it provides values of Pi and Uc for Main Belt
asteroids and Kuiper belt objects which are in agreement
with those previously reported in literature.
The direct source of uncertainty in the computation of
Pi and Uc is due to the casual fluctuations in the random
sampling of the phase space, depending on the total number
of samples (and so the total duration of the computation).
Another, but indirect, error is introduced by the fluctuations
in the construction of the ring model. The list of planetesi-
mal orbits is generated randomly and not all the particles in
the ring intersect the orbit of the target, but only a fraction
of them within a more or less wide range of semimajor axes.
This means that, depending of the details of the random
generation of the orbits, the number of bodies intersecting
the orbit of the target can change a little, impacting on the
evaluation of Pi and Uc. The total number of random sam-
ples has been tuned in order to constrain the uncertainties
within few percents of Pi and Uc.
4 THE "STANDARD" CASE
As a first test, we apply our algorithm to compute the
intrinsic probability of collision and average impact ve-
locity to a "standard" case similar to that explored by
Mustill & Wyatt (2009). In this model, a planetesimal belt
extends from 10 to 20 AU and it is perturbed by a Jupiter --
sized planet with mass mp = 0.001 m⊙ on an orbit with
semimajor axis apl = 5 AU while the eccentricity can be
either epl = 0.1 or epl = 0.6. We consider two distinct con-
figurations: a cold belt that, before the arrival of the planet
in a perturbing orbit, had a very low proper eccentricity
(ep0 ∼ 0). In this case, when the planet turns on its secular
perturbations, the value of ep becomes approximately equal
to ef (see Fig.1). We also explore the evolution of warm
belts that, prior to the planet arrival, had already a high
value of proper eccentricity ep0, possibly excited by other
mechanisms like an extended period of planet -- planet scat-
tering or the formation of large embryos. In this scenario we
assume that ep0 is a significant fraction of ef and we test
three cases, one with ep0 = 0.25 · ef , one with ep0 = 0.5 · ef
and one, the most excited, with ep0 = ef . This is just an
arbitrary choice to show the effects of an initial value of
ep0 > 0 and it is not dictated by any particular scenario. It
appears a better choice than a random selection of ep0 val-
ues. As soon as the planet begins to perturb the warm belt,
the eccentricity values will be encompassed between ef − ep
and ef + ep where ep is derived from the initial value of ep0
as described in Sect. 2 (see Fig.2).
4.1 Initially cold belt
In an initial cold belt the correlation between e and
that the planet establishes when approaching the belt is the
strongest due to the assumption that ep0 = 0. The secular
dynamics is well described by the plots of Fig.1. Both Pi and
Uc are shown in Fig. 4 as a function of the semimajor axis of
the tracers which well approximate the average radial dis-
tance from the star of each body. In this figure we illustrate
the case with epl = 0.1, a configuration in which the plan-
etesimal belt is less perturbed (compared to epl = 0.6). The
green full squares are the impact velocity values computed
assuming there is no secular correlation between the eccen-
tricity and perihelion longitude (hereinafter the 'canonical'
case) while the empty red squares show the probability and
velocity values when the secular correlation is taken into
account. The dotted line in the lower panel of Fig. 4 gives
the average relative velocity computed as Uc = 1.4ef vkep
where vkep is the local Keplerian velocity (Mustill & Wyatt
2009). This relation leads to a simple r−3/2 scaling of the
velocity with the radial distance r. The correlation between
eccentricity and perihelion longitude reduces both Pi and,
in particular, the relative impact velocity Uc respect to the
case without secular correlation. This is due to two effects
which are manifest in Fig. 1: there is a consistent fraction of
orbits with very low eccentricity while the high eccentric or-
bits have their pericentres almost aligned around 0o leading
to low velocity impacts.
The analytic prediction of Mustill & Wyatt (2009) for
Uc appears to slightly overestimate the impact speed at 10
AU by about 10% and underestimate it at 20 AU by approx-
imately the same amount. The r−3/2 scaling does not appear
to fully account for the dynamical evolution of the belt pos-
sibly because it does not properly account for the spatial
distribution of the encounters radial locations related to the
secular dynamics. For Pi, it is difficult to derive a proper
scaling with the radial distance since two different effects
come into play in determining Pi as a function of r. On one
side there is a pure dynamical dependence of the probability
of collision of a planetesimal pair on r related to their or-
bital elements. A typical equation giving such a probability
can be found in Wetherill (1967) (eq. 20). However, we are
computing the Pi not of a single pair, but for an entire pop-
ulation of planetesimals and their radial distribution comes
strongly into play. This radial distribution does not only de-
pend on the planetesimal superficial density distribution but
it is significantly influenced by the distribution of all orbital
elements of the population. As a consequence, it cannot be
predicted a priori with an easy power law distribution, in
particular for highly eccentric belts.
Close to the inner edge of the belt, the value of Pi de-
creases due to the inner truncation of the belt and to a re-
duction in the number of crossing orbits. This effect will dis-
appear for higher eccentricities since the perihelion -- aphelion
Collisional parameters of planetesimal belts
9
3.0
2.0
1.0
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
)
s
/
m
k
(
c
U
1.2
1
0.8
0.6
0.4
0.2
Corr
No-Corr
10
12
14
16
18
20
a (AU)
Corr
No-Corr
10
12
14
16
18
20
a (AU)
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
)
s
/
m
k
(
c
U
10.0
8.0
6.0
4.0
2.0
6.4
5.6
4.8
4
3.2
2.4
1.6
Corr
No-Corr
10
12
14
16
18
20
a (AU)
Corr
No-Corr
10
12
14
16
18
20
a (AU)
Figure 4. Intrinsic probability of collision Pi (top panel) and
impact velocity Uc (bottom panel) for a planetesimal belt per-
turbed by a Jupiter -- size planet with ap = 5 and ep = 0.1. The
belt is assumed to have been initially cold so that ep = ef . The
green filled squares mark the predictions of the canonical model
where no secular correlation is assumed between e and while
the empty red squares are the values computed with the model
that include the secular correlation. The dashed line in the bot-
tom panel outlines the values derived from the analytic formula
Uc = 1.4ef vkep (Mustill & Wyatt 2009).
radial distance is larger for the planetesimals and the orbital
crossing is more extended.
In Fig. 5 we show both Pi and Uc when the eccentricity
of the planet is increased to 0.6, a significantly more per-
turbed configuration for the planetesimal belt. The value of
Pi is increased by about a factor 4 leading to a very ac-
tive belt in terms of collisions while the impact velocity Uc
is five times higher compared to the case with epl = 0.1
and comparable to the value of the present asteroid belt.
The vast majority of collisions are expected to be highly en-
ergetic and both fragmentation and cratering are possibly
dominant leading to a high rate of dust production. These
results show that a higher eccentricity of the planet not only
leads to higher impact speeds but it substantially increases
the impact rate (larger Pi) which is possibly more important
in producing brighter debris disks.
Figure 5. Same as Fig. 4 but for ep = 0.6.
Our numerical algorithm allows to compute the contri-
bution to the average Pi and Uc coming from a restricted
arc of the trajectory of the target, or, more precisely, from
a given range of values of the true anomaly f . Thanks to
this feature, we can evaluate the frequency of collision and
impact velocity around the pericentre and apocentre, respec-
tively. We have selected a range of ±15o around both f = 0o
and f = 180o for each target in order to underline different
values of Pi and Uc in these restricted ranges. The results
are shown in Fig. 6 for the models in which the eccentricity
of the planet is set to ep = 0.1 and ep = 0.6, respectively.
When ep = 0.1, the difference between Pi and Uc at
pericentre and apocentre is small. However, when ep = 0.6
a significantly higher value of Pi is observed when the target
is at the apocentre and hence in the inner regions of the belt.
Beyond 17 AU the trend reverses and Pi becomes larger at
pericentre. This behaviour is due to the high forced (and
then proper) eccentricity induced by the secular perturba-
tions of the planet. When the target is close to the inner
border of the belt its pericentre is located where the density
of potential projectiles is significantly lower due to the over-
all eccentricity distribution ranging from 0 and 2 ef . On the
other hand, when the target orbits around the apocentre, it
10
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
)
s
/
m
k
(
c
U
1.5
1.0
0.5
0.0
8
7.2
6.4
5.6
4.8
4
3.2
2.4
1.6
0.8
0
ep = 0.1, per
ep = 0.1, apo
ep = 0.6, per
ep = 0.6, apo
10
12
14
16
18
20
a (AU)
ep = 0.1, per
ep = 0.1, apo
ep = 0.6, per
ep = 0.6, apo
10
12
14
16
18
20
a (AU)
Figure 6. Intrinsic probability of collision Pi (top panel) and im-
pact velocity Uc (bottom panel) computed around the pericentre
(−15o < f < 15o) and apocentre (165o < f < 195o) of all targets.
The eccentricity of the planet is set to ep = 0.1 and ep = 0.6.
is well within the belt where the density of the projectiles is
the highest and Pi is large. Moving towards the outer bor-
der of the belt, it is now the apocentre of the targets that
is located in low density regions while the pericentre is well
within the belt and this leads to higher values of Pi. Super-
imposed to this effect there is also the decreasing trend of Pi
for larger values of a due to a reduction of both the forced
eccentricity and radial density of the planetesimal popula-
tion. The values of Pi shown in Fig. 6 are only a fraction
of the total impact probability illustrated in Fig. 5 since
we are considering only a portion of the total range of the
true anomaly f and, as a consequence, they are substantially
smaller than the values in Fig. 5.
Significantly different values of Uc are also found at peri-
centre and apocentre when ep = 0.6. At apocentre the values
of the impact speed are comparable with the average values
shown in Fig. 5 but at pericetre Uc is much higher since
the target has a higher orbital velocity compared to the po-
tential projectiles. Due to the correlation between e and ,
a significant number of impacts when the target is at the
pericentre occurs with circular orbits having radius approx-
imately equal to a(1 − e). By comparison, at apocentre the
target will frequently encounter projectiles on circular or-
bits with radius a(1 + e). As a consequence, at pericentre
the impact velocity Uc is higher.
4.2 Initially warm belt
In a scenario where the planet is injected in an orbit perturb-
ing an already heated planetesimal belt, the values of Pi and
Uc are different respect to those of a cold belt because of the
changes in the secular dynamics (Fig. 2). Here we consider
three test cases where the proper eccentricity ep0, before the
onset of the planet perturbations, is 0.25, 0.5 and 1.0 that
of the forced eccentricity (the value that will be established
once the planet approaches the belt). Different values of ep0
can be encountered in real systems and our choice to link ep
to ef is arbitrary, but it is impossible to explore all possible
initial distributions of ep0, so we consider only three refer-
ence cases, where ep0 is linearly related to ef , which give
clues on how to deal with any general scenario.
In Fig. 7 we compare Pi and Uc when epl = 0.1 for
all the different initial configurations corresponding to dis-
tinct dynamical states of the belt prior to the onset of the
planet perturbations, i.e. cold and warm belts. This compar-
ison highlights the effects on Pi and Uc of different degrees
of correlation between e and related to the distinct ini-
tial values of ep0. Both the intrinsic probability of collision
and impact velocity increase for higher values of ep0 (ini-
tially warm belts) until, for ep0 = ef , values comparable
to those of the uncorrelated case (the canonical one) are
obtained. This increase is due to the complex interplay of
dynamical effects which can be deduced from Fig.2. First
of all in warm belts higher eccentricity values are achieved
since ep ranges from ef − ep0 to ef + ep0 respect to the cold
scenario where ep = ef . In addition, when the pericentre
is around 180o the planetesimal eccentricities are not all
small as illustrated in Fig.1 but they may reach significant
values (Fig.2). Finally, pseudo -- librators, whose fraction de-
pends on the initial value of ep0, impact with non -- librators
leading to geometrical configurations where the trajectories
at the crossing are more bent and favour larger collisional
velocities. All these mechanisms contribute to increase both
Pi and Uc up to values similar to those of the canonical case
also when epl = 0.6, as illustrated in Fig. 8.
Our numerical models show that initial warm belts have
a more intense collisional activity due to higher values of
both Pi and Uc. This leads to a greater production of dust
and a brighter debris disk associated to the belt but also to
a faster erosion rate with a shorter lifetime.
The value of impact velocity Uc shown in the previous
plots is indeed an average over a large number of computed
impact speeds. The velocity distribution in the standard case
with epl = 0.6 for the target with aT = 15 AU is displayed
in Fig. 9. The black dashed line shows the impact velocity
distribution in absence of e -- correlation (a warm belt not
perturbed by a planet) while the red continuous line illus-
trates the distribution when such correlation is included in
the computation of Uc for an initially cold belt. The high
velocity tail of the distribution is cut off and this explains
the reduction in the average impact speed observed in Fig.5.
In the warm case with ep0 = 0.5ef the cut off at high impact
speeds is reduced compared to the case with ep0 = 0 due to
the contribution of pseudo -- librators impacting circulators.
3.0
2.0
1.0
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
1
0.8
0.6
0.4
)
s
/
m
k
(
c
U
Collisional parameters of planetesimal belts
11
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = 1.0 ef
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
8.0
6.0
4.0
2.0
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = 1.0 ef
10
12
14
16
18
20
10
12
14
16
18
20
a (AU)
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = 1.0 ef
6.4
5.6
4.8
4
3.2
2.4
)
s
/
m
k
(
c
U
a (AU)
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = ef
10
12
14
16
18
20
10
12
14
16
18
20
a (AU)
a (AU)
Figure 7. Pi (upper panel) and Uc (lower panel) for initially
warm and cold belts. For the warm belts we assume three different
values of ep0, i.e. 0.25 (green empty squares), 0.5 (blue empty
squares) and 1 times ef (magenta empty squares). The red filled
squares mark the values of Pi and Uc for a cold belt while the
black empty circles show the case with ep = ef without secular
perturbations. The eccentricity of the perturbing planet is set to
epl = 0.1.
However, the peak is located at lower impact speeds and
the average value of Uc is still lower compared to the case
without e -- correlation.
n
o
i
t
c
a
r
F
4.3 Initially warm and inclined belt: i = 15o
In presence of a strong dynamical excitation, due to possi-
ble different mechanisms like self -- stirring, the presence of
large planetary embryos, resonance sweeping etc., even the
osculating inclinations may be significantly pumped up. We
explore in this section the effects of increasing the average
value of inclination on the values of Pi and Uc. Intuitively,
one would expect an increase in the relative impact velocity
due to the presence of an additional out -- of -- plane compo-
nent in the difference between the velocity vectors of two
planetesimals. At the same time, due to an expansion of the
available space for orbital motion, a significant decrease of
the impact probability is expected. Both these predictions
Figure 8. Same as Fig. 7 but for epl = 0.6.
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
ep0 = 0 No-Corr
ep0 = 0 Corr
ep0 = ef/2 Corr
0 1 2 3 4 5 6 7 8 9
v (km/s)
Figure 9. Normalised impact velocity distribution for ep = ef
without the secular correlation between e and (black dashed
line) and for the case in which such correlation is included in the
computations (red continuous line). The blue dotted line illus-
trates the velocity distribution for a warm belt with ep0 = 0.5· ef .
0.25 -- 3o
0.5 -- 3o
1.0 -- 3o
0.25 -- 15o
0.5 -- 15o
1.0 -- 15o
10
12
14
16
18
20
a (AU)
0.25 -- 3o
0.5 -- 3o
1.0 -- 3o
0.25 -- 15o
0.5 -- 15o
1.0 -- 15o
12
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
8.0
6.0
4.0
2.0
6.4
5.6
4.8
4
3.2
2.4
)
s
/
m
k
(
c
U
10
12
14
16
18
20
a (AU)
Figure 10. Intrinsic probability of collision Pi and impact ve-
locity Uc for a small body belt with ep0 = 0.25, 0.5 and 1.0 ef
and i = 15o, compared to the values computed for i = 3o. The
perturbing planet has an eccentricity epl = 0.6. While Pi is signifi-
cantly reduced, the impact velocity Uc is approximately increased
by 40%.
are confirmed by the outcome of our algorithm and in Fig. 10
we show the computed values of Pi and Uc for initially warm
belts with ep0 = 0.25, 0.5, 1 · ef and an average inclination
of 15o (the perturbing planet has epl = 0.6). The explo-
ration of larger values of inclination would require a more
detailed secular approach where the inclination is coupled
to the node longitude and this will be done in a forthcoming
paper.
In the case of inclined belts, the value of Pi in the inner
regions of the belt is reduced by about a factor of 4 compared
to the case with i = 3o (see Fig.10). On the other hand, the
impact velocity Uc is higher for more inclined planetesimals
due to the out -- of -- plane component in the relative velocity,
reaching a maximum value of 8 km/s at 10 AU from the star
when ep0 = 1 · ef (magenta empty squares in the bottom
panel of Fig. 10). However, a higher inclination appears to
be not as important as a higher values of ep0 in leading to
higher impact speed. In general, according to our modelling,
an inclined belt is expected to be less collisionally active due
to the strong decrease in Pi, in spite of an increase of Uc.
This will be confirmed by the collisional evolution models of
the next section.
4.4 Collisionally damped belt
We consider here the case of a densely populated belt whose
proper eccentricity distribution, after the approach of the
planet, has been collisionally damped. In this scenario most
of the orbital excitation is dissipated and the proper ec-
centricity is reduced to values smaller than the forced one.
This would lead to the dynamical state illustrated in Fig.3
where all the planetesimals are in pseudo -- libration. We com-
pute the intrinsic probability of collision and impact veloc-
ity when the damped configuration is reached. To model this
kind of belt we select a value of proper eccentricity, after the
damping, which is half the value of the forced one. Even in
this case, the choice is made to give an idea of the influence
of this secular configuration on the collisional evolution of
the belt, i.e. on the values of Pi and Uc, and it is not related
to any particular scenario.
For a damped belt both Pi and Uc are reduced (see
Fig.11) and this has to be ascribed to the pseudo -- libration of
all planetesimals. oscillates within a limited range around
0o leading to a significant level of pericentre alignment of
all orbits. This dynamical configuration leads to collisions
where the trajectories of the approaching planetesimals are
almost parallel at the orbital crossing. As a consequence,
there is a consistent reduction of the impact velocity and
impact rate. A damped belt is then expected to be less col-
lisionally active and it will give rise to debris disks which
are less bright.
5 COLLISIONAL EVOLUTION
collisional
To explore the implications of different values of Pi
and Uc on the evolution of a planetesimal belt we
have run a simple one-dimensional
evolu-
tion code (Campo Bagatin et al. 1994; Marzari et al. 1995;
Bottke et al. 2005). We start from an initial planetesimal
population extending from 10 to 500 km in diameter and
distributed in a series of discrete logarithmic diameter bins
following a slope equal to -3.5. The population is centred at
15 AU and it is 10 AU wide. The code computes the col-
lisional interactions of each size bin with every other one
during a given time -- step. At the end of the time -- step, all
the outcomes of the interactions are summed up to find the
net change in the population as a function of size. The up-
dated population is then used in the next time -- step until the
whole time -- span, that we assume to be 4 Gyr, is covered.
These calculations are performed under the assumption that
the orbital element distribution of the planetesimals is not
significantly altered by the collisions so that both Pi and Uc
remain constant during the entire evolution of the belt. Two
different scaling laws for the computation of the fragments
size distribution after a catastrophic impact are tested: a
simple energy scaling like that proposed in Davis & Farinella
(1997) to study the collisional evolution of Kuiper Belt ob-
jects and the more recent one given in Stewart & Leinhardt
(2009) for weak aggregates (their Eq. 2). The results are not
significantly different so we report only those obtained with
the scaling law of Stewart & Leinhardt (2009).
8.0
6.0
4.0
2.0
0
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
)
s
/
m
k
(
c
U
4.8
4
3.2
2.4
1.6
0.8
10
12
14
16
18
20
a (AU)
ep = ef/2
ep = ef
10
12
14
16
18
20
a (AU)
Figure 11. Intrinsic probability of collision Pi and impact ve-
locity Uc for a small body belt collisionally damped after the
onset of the planet perturbations. The eccentricity of the planet
is set to 0.6 and the proper eccentricity is assumed to have been
damped to a value ep = 0.5 · ef . Both Pi and Uc are compared
to the values computed for a non -- damped cold belt (filled green
squares). There is a significant decrease of both Pi and Uc due to
the pseudo -- libration of all planetesimals.
We have run four different models: two for cold belts
with epl = 0.1 and epl = 0.6, one for an initially warm belt
where epl = 0.6 and ep0 = 0.5 · ef , and the last case for
epl = 0.6, ep0 = 0.5 · ef and i = 15o. The evolved planetesi-
mal populations in the four different cases are shown in Fig.
12. When epl = 0.6, the erosion of the belt leads to a signifi-
cant reduction in the number of planetesimals even at large
sizes. Compared to the case with epl = 0.1, the simulation
with epl = 0.6 shows an erosion rate about 10 times faster
over 4 Gyr. A significant larger amount of dust is then ex-
pected to be produced when the eccentricity of the planet
is higher, at least in the initial phases of evolution of the
belt, as it could have been argued by the larger values of
both Pi and Uc. However, old belts may have been signifi-
cantly depleted by the the initial fast collisional erosion and
show at present a dust production rate comparable or even
lower than that of young and less depleted belts around low
eccentricity planets.
The collisional evolution of an initial warm belt with
Collisional parameters of planetesimal belts
13
ep = ef/2
ep=ef
1e+10
1e+08
1e+06
10000
100
n
b
i
r
e
t
e
m
a
d
i
r
e
p
s
e
d
o
b
i
f
o
.
N
IP
0.1c
0.6c
0.6w
0.6i
1
0.1
1
10
D(KM)
100
1000
Figure 12. Collisional evolution over 4 Gyr of a putative plan-
etesimal belt, centred at 15 AU and 5 AU wide, for two different
values of the planet eccentricity epl, i.e. 0.1 and 0.6, and for a
warm and inclined population. The values of Pi and Uc used in
the collisional evolution code are taken from the previous compu-
tations. The initial size distribution (red line), divided in a series
of discrete bins, is truncated at 10 and 500 km in diameters. The
erosion of the belt is faster in the case labelled 0.6w (magenta
circles) where epl = 0.6 and ep0 = 0.5 · ef , i.e. a warm belt. The
initial cold belt with the same planet eccentricity (0.6c) (blue
filled circles) differs only slightly from the 0.6w case. When the
inclination of the warm belt is increased to 15o the collisional
evolution is reduced, as expected by the strong decrease of Pi
(see Fig.10). The case with epl = 0.1, labelled 0.1c, illustrates
the evolution of a cold belt perturbed by a low inclination planet
and it is the less collisionally active case.
ep0 = 0.5 · ef and epl = 0.6 is only slightly faster compared
to an initial cold belt (ep0 = 0) and a similar dust production
rate is expected. This shows that, in spite of an increase of
the impact velocity Uc in the warm case, the two dynamical
configurations lead to a similar collisional evolution because
Pi is almost equal. This is confirmed also by the inclined case
where a lower value of Pi leads to a less eroded belt in spite of
a higher collisional velocity (see Fig.10). This suggests that
the intrinsic probability of collision is decisive in determining
the erosion of a small body belt notwithstanding different
values of impact velocity.
The previous modelling must be considered only as in-
dicative since additional information are needed in particu-
lar about the structural strength and porosity of the plan-
etesimals since these physical aspects are heavily involved
in the computation of the outcome of mutual collisions. In
addition, this code is limited in the size range that it can
cover and it does not possess spatial resolution. It cannot be
compared with codes like DyCoSS and LIDT-DD (Th´ebault
2012; Kral et al. 2013) describing the evolution of a de-
bris disk both in time and space, including effects like the
Poynting -- Robertson drag and creating a coupling between
dynamics and collisions. The only advantage of our simpli-
fied code is that it can predict on the long term (timescale of
the order of Gyrs) how the dust production rate will decline
and establish a potential correlation between the age of the
star and the luminosity of the belt. It can also be used in a
reverse approach allowing to determine the size distribution
14
of the primordial belt from the present belt around an aged
star, as it is usually done for the asteroid belt in the solar
system.
6 THE BELT IN HD 38529: COMPARISON
WITH ǫ ERIDANI
HD 38529 is a complex system where two massive plan-
ets have been found to orbit the central body, a 3.5 Gyr
old G8 type star. Their masses are 0.8 and 12.2 MJ and
their orbital elements are a1 = 0.13 AU , a2 = 3.74 AU,
e1 = 0.25 and e2 = 0.35, respectively (Butler et al. 2006).
Moro-Mart´ın et al. (2007b) found from Spitzer data an in-
frared excess that interpreted as due to the presence of a
debris disk. From dynamical constraints they located the
dust -- producing long -- lived planetesimals in three regions. A
small inner ring orbiting between the two planets from 0.4 to
0.8 AU, a wider outer belt extending from 20 to 50 AU and
an even outer one beyond 60 AU. The two outer belts are
separated from a strong secular resonance located at about
55 AU. We focus our study on the belt extending from 20 to
50 AU where Moro-Mart´ın et al. (2007b) constrain the loca-
tion of the dust -- producing planetesimals. They in fact find
that the debris disk is collision -- dominated implying that
the planetesimals approximately share the same orbits as
the emitting dust. We compute for this system the intrin-
sic probability of collision and impact velocity at different
radial distances within 20 and 50 AU. In Fig. 13 we show
Pi and Uc for this belt. We consider only the outer more
massive planet as perturber and we neglect the influence of
the smaller inner one. Compared to the Kuiper Belt in the
solar system, the intrinsic probability of collision appears
to be larger in particular in the inner regions of the belt.
According to Dell'Oro et al. (2001), Pi in the Kuiper Belt
ranges from 3.14 to 4.44 ×10−22 yr−1 km−2 while from Fig.
13 for 38529 Pi is at least three times larger in between 20
to 30 AU. On the other hand, the impact velocity in the
HD 38529 debris disk appears to be smaller than that in the
Kuiper belt whose value is about 1.23 -- 1.44 km/s. In short,
in HD 38529 collisions are expected to be more frequent but
less energetic.
By using the values of Pi and Uc at the centre of the
belt (35 AU) (see Fig. 13) we run the collisional evolution
code. In Fig. 14 the evolution with time of the size distri-
bution of an initial planetesimal belt equal to that used for
Fig. 12 is shown. Compared to Fig. 12, the shattering events
are significantly reduced and only a limited amount of bod-
ies smaller than 100 km in diameter are catastrophically
disrupted. This must be ascribed to the lower value of Pi
which is about 30 times smaller for the HD 38529 belt com-
pared to the Pi of our standard case. The dust production
rate is then expected to be significantly lower. The difference
observed between the cold and warm case at small sizes is
due to the slightly larger number of big bodies disrupted
in the warm belt, phenomenon which can barely be seen in
Fig. 12 in between 10 and 40 km. The fragments produced
in these breakups refill the size distribution at smaller di-
ameters explaining the larger number of small bodies in the
warm belt.
Together with HD 38529, also ǫ Eridani has a de-
bris disk identified by dust emission (Greaves et al. 1998;
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = 1.0 ef
ep = 0.5 ef
20
25
30
35
40
45
50
a (AU)
ep0 = 0
ep0 = 0.25 ef
ep0 = 0.5 ef
ep0 = 1.0 ef
ep = 0.5 ef
2
2
-
0
1
x
)
1
-
r
y
2
-
m
k
(
i
P
)
s
/
m
k
(
c
U
18.0
15.0
12.0
9.0
6.0
3.0
1
0.8
0.6
0.4
0.2
20
25
30
35
40
45
50
a (AU)
Figure 13. Pi (upper panel) and Uc (lower panel) for the belt
observed in the HD 38529 system. We consider three different
models: an initial cold belt with ep = ef (ep0 = 0), an initial
warm belt with different values of ep0 and a collisionally damped
belt with ep = 0.5 · ef .
MacGregor et al. 2015). The belt is supposed to extend from
approximately 53 to 80 AU from the central star and it is
perturbed by a giant planet with a mass approximately 1.66
times the mass of Jupiter and on an orbit with semimajor
axis equal to 3.39 AU and eccentricity of 0.7 (Benedict et al.
2006b). This belt has approximately the same radial exten-
sion of that around HD 38529 (about 30 AU) but it is located
farther out from the star. The mass in the dust estimated in
the ǫ Eridani debris disk is about 2 × 10−5M⊙, according to
Li et al. (2003). This value is significantly higher compared
to that inferred for HD 38529 by Moro-Mart´ın et al. (2007b)
of 1−5×10−10 even if this was obtained for particles of 10 µm
in size. A higher dust mass for ǫ Eridani can only partially be
explained by the younger age of the system (about 800 Myr
for ǫ Eridani vs. 3.28 Gyr for HD 38529). We have computed
the intrinsic probability of collision and impact velocity for
the putative planetesimal belt precursor of the debris disk
in ǫ Eridani. At the centre of the belt around 67 AU we find
Pi ∼ 2 × 10−23 km−2yr−1 and an average impact velocity
of about Uc ∼ 0.1km/s while for HD 38529 in the centre of
the belt at 35 AU (see Fig. 13) we have Pi ∼ 4.5 × 10−22
Initial pop.
ep0 = 0
ep0 = 0.5 ef
1e+10
1e+08
1e+06
10000
100
n
b
i
r
e
t
e
m
a
d
i
r
e
p
s
e
d
o
b
i
f
o
.
N
1
0.1
1
10
D(KM)
100
1000
Figure 14. Collisional evolution of a planetesimal belt in the HD
38529. The erosion rate is much slower compared to that of the
standard case shown in Fig. 12 due to the lower values of both Pi
and Uc. The warm belt is slightly collisionally more active since a
larger number of bodies are disrupted in between 10 and 40 km in
diameter. These breakups produce smaller fragments which refill
the population at small sizes.
km−2yr−1 and Uc ∼ 0.35km/s (assuming in both cases an
initial cold belt). The collisional activity is then expected to
be at least 20 times faster in HD 38529 respect to ǫ Eridani.
However, as shown in Fig.14, the collisional evolution in HD
38529 cannot have led to a significant erosion of the belt
even after 3.28 Gyr. Why is then the dust in the debris disk
around ǫ Eridani more dense than that around HD 38529?
The only possible explanation is that the planetesimal belt
in ǫ Eridani is significantly more populated respect to that
of HD 38529. The star of ǫ Eridani is less massive compared
to that of HD 38529 (0.83 vs. 1.39 M⊙) so it is difficult to
imagine that the protoplanetary disk around ǫ Eridani was
much more dense than that around HD 38529. One possibil-
ity is that the dynamical history of HD 38529 has been more
turbulent with extended periods of planet -- planet scattering.
This behaviour might have depleted the belt, as suggested
by (Bonsor et al. 2013; Marzari 2014), leaving a lighter belt
that subsequently evolved under mutual collisions. This evo-
lution might explain the present difference between the two
systems.
7 DISCUSSION AND CONCLUSIONS
The dust production rate in a debris disk, and then its
brightness, strongly depends on the collisional evolution of
the parent planetesimal belt. Since planetesimals are rem-
nants of the planet formation process it is then reasonable
to expect that one or more planets formed in the system,
notwithstanding any significant correlation between stars
with dust emission and the presence of known planets is still
under scrutiny. The collisional evolution of planetesimals
and the dust production rate can be determined once both
the physical structure of the bodies (in particular strength
and porosity) and two fundamental dynamical parameters,
the intrinsic probability of collision Pi and the mutual im-
pact velocity Uc, are known. While the physical structure
Collisional parameters of planetesimal belts
15
of planetesimals depends on the dust accumulation process
in the early phases of the circumstellar disk evolution, the
values of the two parameters Pi and Uc are uniquely de-
termined by the orbital architecture of the system. This
in turn depends on planetesimal accumulation and planet
formation but it is possibly influenced also by other evolu-
tionary processes like planet migration by tidal interaction
with the gaseous disk and phases of planet -- planet scatter-
ing. This last mechanism is expected to have occurred in a
significant fraction of systems due to the large number of
exoplanets found in highly eccentric orbits. Planet -- planet
scattering not only excites eccentricity but it also moves
planets around during the chaotic period characterised by
mutual close encounters between the planets prior to the
ejection of one (or more) out of the system.
A planetesimal belt can be affected by the dynami-
cal behaviour of the planets and we can envision different
scenarios in which planets and planetesimal belts interact,
evolve and finally coexist. We are interested in a scenario
where a planet perturbs the belt from an inside eccentric
orbit. This configuration can be achieved via a smooth path
where the planet, after its formation, evolved close to the
belt without dramatic dynamical events and in a low ec-
centricity orbit. In this case the belt is expected to be ini-
tially cold before the onset of the secular perturbations of
the planet. If instead the planet is involved in a period of
planet -- planet scattering, its present location close to the
belt would be the outcome of a chaotic past and its eccen-
tricity is expected to be high. In this second scenario, the
belt may have been affected by a period of planet -- planet
scattering at different levels. It may have been spared by
most of the chaotic evolution of the planets and, when one
planet ends up in a close orbit, be directly affected by its sec-
ular perturbations starting from a non -- excited state (initial
cold belt). In alternative, it may have been stirred up by the
planets evolving on highly eccentric orbits induced by the
mutual close encounters before a quiet state is established
following the ejection of one or more bodies. In this case a
warm belt with excited eccentricities and, possibly, inclina-
tions would interact with the planet which finally ends up
in an orbit close to the belt at the end of the chaotic phase.
Intermediate states can also be envisioned where the belt
is warmed either by self -- stirring induced by the formation
of large embryos or by resonance sweeping by a migrating
planet or other possible mechanisms before being perturbed
by an incoming planet.
In this paper we apply a semi-analytic method specifi-
cally designed to compute both Pi and Uc for belts perturbed
by a nearby planet. Its secular perturbations force a strong
correlation between the eccentricity e and perihelion longi-
tude of the planetesimal orbits significantly affecting the
values of both Pi and Uc. The knowledge of these two dy-
namical parameters allow to model the collisional evolution
of a belt and quantitatively predict its erosion over the star
age. With suited collisional models these parameters can also
be used to infer the early planetesimal population that led
to the present belt whose properties are deduced by the re-
lated debris disk. From an observational point of view, they
can be used to relate the star age to the brightness of the
debris disk, taking of course into account the uncertainties
on the initial planetesimal population, its physical proper-
ties and the planet orbital parameters. It can also predict
16
the frequency of large breakup events in the belt which can
lead to a sudden significant enhancement of the brightness
of the related debris disk. The parameters we compute can
also be used in more refined codes predicting the luminosity
of a debris disk for a given planetesimal belt.
We consider distinct dynamical configurations in which
the planet perturbs either an initial cold or warm belt. In
the case of a warm belt, we also contemplate the case in
which there is a significant inclination excitation. In addi-
tion, we compute the values of Pi and Uc for crowded belts
where a significant collisional damping may occur. We focus
on a standard case with a Jupiter -- sized planet orbiting at
about 5 AU from the star and a planetesimal belt extending
beyond 10 AU. We show that the secular perturbations of
the planet and, in particular, the correlation between e and
typical of these perturbations, cause a reduction of both
Pi and Uc respect to a non -- correlated case. This last case
may occur, in absence of a close perturbing planet, if the
planetesimal belt self -- stirred for example via the formation
of large planetary embryos, or if it was perturbed by planets
roaming around during a chaotic phase and, when finally a
quiet state is reached, it is located far away from any po-
tentially perturbing bodies. This finding suggests that the
presence of a perturbing planet not always leads to a higher
dust production rate and that self -- stirred planetesimal belts
might give origin to brighter debris disks. We have tested dif-
ferent values of the planet eccentricity finding higher values
of both Pi and Uc in presence of more eccentric planets.
We also explore as Pi and Uc change when a belt was
warm before being secularly perturbed by a planet. In this
case, the approximation ep = ef cannot be adopted and
the proper eccentricity, after the onset of the planet per-
turbations, can assume a range of values from ef − ep0 to
ef + ep0 where ep0 is the average eccentricity of the warm
planetesimal belt before the planet begins to perturb it. In
our modelling we assume that ep0 is a fixed fraction of the
forced eccentricity, an arbitrary choice simply dictated by
the need of avoiding a random selection of initial values. We
do not consider scenarios where ep0 is initially larger than ef .
For warm belts we find an increase in the collisional activity
respect to initially cold belts but they are still less colli-
sionally eroded compared to the 'self -- stirred' belts, at least
until ep0 6 ef . A warm belt is then expected to be gradually
depleted on shorter timescales. A dynamically excited belt
may also have high inclinations. In this configuration an in-
crease of the relative velocity and a decrease of the intrinsic
probability of collision are found. However, the two effects
do not compensate and the reduction of Pi leads to a lower
erosion rate compared to a small body belt with low inclina-
tion. This proves the dominating role of Pi in determining
the collisional evolution of a minor body population.
We have applied our formalism to model the case of
the debris disk detected in HD 38529. The collisional activ-
ity in the putative belt hosting the leftover planetesimals is
low compared to the standard case we have studied in this
paper and after 4 Gyr of collisional evolution only a low
percentage of bodies smaller than 100 km are expected to
have been disrupted by collisions. However, the planetesi-
mal belt appears to be more collisionally active compared
to that detected around ǫ Eridani. This finding is at odds
with the observations of the debris disks around the two
stars with that around ǫ Eridani being more dense respect
to that around HD 38529. One possible explanation is that
a turbulent dynamical past depleted the belt around HD
38529 reducing the number density of planetesimals.
On the basis of the secular theory, we can envisage the
following evolution of the brightness of a debris disk associ-
ated with an initially cold belt. At the beginning, when the
planet approaches the belt, the dust production rate will
slowly increase during the progressive randomisation of the
pericentre. When the s are still in phase, low impact veloc-
ities are expected but at subsequent times the progressively
different values of pericentre longitudes will lead to more
energetic collisions and a higher dust production rate. The
debris disk will become brighter. When the full randomisa-
tion is reached, from then on the dominant mechanism will
be the slow erosion of the planetesimal belt due to mutual
collisions and the brightness of the associated debris disk
will slowly decrease with time. The rate of luminosity de-
crease can be estimated on the basis of Pi and Uc which are
the basic parameters to determine the collisional evolution
of the belt.
The situation appears slightly more complex for an ini-
tially warm belt. At the beginning, just after the arrival of
the planet, the pericentre evolution quickly leads pseudo --
librators to cross the orbits of circulators causing a fast rise
in the collisional rate and an increase in the luminosity of
the associated debris disk. However, at subsequent time this
initial highly collisional state will develop into a dynamical
configuration where the the secular perturbations are fully
developed and the erosion with follow the same path as in
a cold belt even if with different speed due to the higher
values of Pi and Uc.
In presence of a strong collisional damping, like in a very
crowded belt, the dynamics is dominated by the pseudo --
libration of all planetesimals and the collisional activity is
strongly reduced with values of both Pi and Uc significantly
smaller respect to both the cases of cold and warm non --
damped belts.
It is clear that, on the basis of the values of Pi and Uc
we can derive different relations between the brightness of a
disk and the age of the star, since the erosion of the belt will
be faster or slower depending on the orbital parameters of
the planet and the dynamical state of the belt at the onset
of the secular perturbations.
In this paper we have focused on the perturbations of
a single planet on a planetesimal belt, but systems of mul-
tiple giant planets perturbing coexisting belts can be found
as well. In this last case a more complex secular evolu-
tion is expected. However, our algorithm for computing Pi
and Uc can be easily applied to these systems once the the
frequencies of perihelion precession of all planets are com-
puted. If the eccentricities of the planets are high, the linear
secular theory may be a rough approximation for comput-
ing the real frequencies and higher order theories may be
needed (Michtchenko & Malhotra 2004; Libert & Henrard
2005). This problem will be faced in a subsequent paper.
Concerning the contribution from mean motion reso-
nances to the collisional evolution of a belt, it can be con-
sidered negligible in a configuration where a single planet
perturbs a wide belt considering the tiny resonance width
even for highly eccentric planets. Together with secular reso-
nances, mean motion resonances may cause local differences
in the brightness of the debris disk associated to the belt
but they are not expected to cause significant differences
in the overall collisional evolution of wide belts. For narrow
belts a dedicated dynamical exploration may be necessary
to evaluate the contribution of resonances.
ACKNOWLEDGMENTS
We thank an anonymous referee for his/her useful comments
and suggestions that helped to significantly improve the pa-
per.
REFERENCES
Collisional parameters of planetesimal belts
17
Kral, Q., Th´ebault, P., & Charnoz, S. 2013, A&A, 558,
A121
Krivov, A. V. 2010, Research in Astronomy and Astro-
physics, 10, 383
Li, A., Lunine, J. I., & Bendo, G. J. 2003, ApJL, 598, L51
Libert, A.-S. & Henrard, J. 2005, Celestial Mechanics and
Dynamical Astronomy, 93, 187
MacGregor, M. A., Wilner, D. J., Andrews, S. M., Lestrade,
J.-F., & Maddison, S. 2015, APJ, 809, 47
Marzari, F. 2014, MNRAS, 444, 1419
Marzari, F., Davis, D., & Vanzani, V. 1995, Icarus, 113,
168
Matthews, B. C., Krivov, A. V., Wyatt, M. C., Bryden, G.,
& Eiroa, C. 2014, Protostars and Planets VI, 521
Meshkat, T., Bailey, V. P., Su, K. Y. L., et al. 2015, ApJ,
Benedict, G. F., McArthur, B. E., Gatewood, G., et al.
800, 5
2006a, Aj, 132, 2206
Benedict, G. F., McArthur, B. E., Gatewood, G., et al.
2006b, AJ, 132, 2206
Michtchenko, T. A. & Malhotra, R. 2004, Icarus, 168, 237
Moro-Mart´ın, A., Carpenter, J. M., Meyer, M. R., et al.
2007a, ApJ, 658, 1312
Bonsor, A., Raymond, S. N., & Augereau, J.-C. 2013, MN-
Moro-Mart´ın, A., Malhotra, R., Carpenter, J. M., et al.
RAS, 433, 2938
Bottke, W. F., Broz, M., B'Brien, D. P., et al. 2002, The
Collisional Evolution of the Main Asteroid Belt, 701 -- 725
Bottke, W. F., Durda, D. D., Nesvorn´y, D., et al. 2005,
Icarus, 175, 111
2007b, Apj, 668, 1165
Moro-Mart´ın, A., Wolf, S., & Malhotra, R. 2005, Apj, 621,
1079
Murray, C. D. & Dermott, S. F. 1999, Solar system dy-
namics
Bottke, W. F., Nolan, M. C., Greenberg, R., & Kolvoord,
R. A. 1994, Icarus, 107, 255
Mustill, A. J. & Wyatt, M. C. 2009, MNRAS, 399, 1403
Opik, E. J. 1951, Proc. R. Irish Acad. Sect. A, vol. 54,
Bryden, G., Beichman, C. A., Carpenter, J. M., et al. 2009,
p. 165-199 (1951)., 54, 165
Apj, 705, 1226
Shannon, A., Bonsor, A., Kral, Q., & Matthews, E. 2016,
Butler, R. P., Wright, J. T., Marcy, G. W., et al. 2006, Apj,
MNRAS, 462, L116
646, 505
Sibthorpe, B., Ivison, R. J., Massey, R. J., et al. 2013,
Campo Bagatin, A., Cellino, A., Davis, D. R., Farinella, P.,
MNRAS, 428, L6
Stewart, G. R. & Wetherill, G. W. 1988, Icarus, 74, 542
Stewart, S. T. & Leinhardt, Z. M. 2009, ApjL, 691, L133
Su, K. Y. L. & Rieke, G. H. 2014, in IAU Symposium, Vol.
299, Exploring the Formation and Evolution of Planetary
Systems, ed. M. Booth, B. C. Matthews, & J. R. Graham,
318 -- 321
Th´ebault, P. 2012, A&A, 537, A65
Th´ebault, P., Marzari, F., & Scholl, H. 2006, Icarus, 183,
193
Trilling, D. E., Bryden, G., Beichman, C. A., et al. 2008,
Apj, 674, 1086
Wetherill, G. W. 1967, JGR, 72, 2429
Whitmire, D. P., Matese, J. J., Criswell, L., & Mikkola, S.
1998, Icarus, 132, 196
Wyatt, M. C., Smith, R., Greaves, J. S., et al. 2007, Apj,
658, 569
This paper has been typeset from a TEX/ LATEX file prepared
by the author.
& Paolicchi, P. 1994, PSS, 42, 1079
Chambers, J. E. & Wetherill, G. W. 2001, Meteoritics and
Planetary Science, 36, 381
Davis, D., Durda, D., Marzari, F., Campo Bagatin, A., &
Gil-Hutton, R. 2002, Collisional Evolution of Small-Body
Populations, ed. W. F. Bottke, Jr., A. Cellino, P. Paolic-
chi, & R. P. Binzel, 545 -- 558
Davis, D. R. & Farinella, P. 1997, Icarus, 125, 50
Dell'Oro, A. 2016, MNRAS, submitted, 183
Dell'Oro, A., Marzari, F., Paolicchi, P., & Vanzani, V. 2001,
A&A, 366, 1053
Dell'Oro, A., Paolicchi, F., Marzari, P., Dotto, E., & Van-
zani, V. 1998, A&A, 339, 272
Dell'Oro, A. & Paolicchi, P. 1997, PSS, 45, 779
Dell'Oro, A. & Paolicchi, P. 1998, Icarus, 136, 328
Everhart, E. 1985,
in Dynamics of Comets: Their Ori-
gin and Evolution, Proceedings of IAU Colloq. 83, held
in Rome, Italy, June 11-15, 1984. Edited by Andrea
Carusi and Giovanni B. Valsecchi. Dordrecht: Reidel, As-
trophysics and Space Science Library. Volume 115, 1985,
p.185, ed. A. Carusi & G. B. Valsecchi, 185
Farinella, P. & Davis, D. R. 1992, Icarus, 97, 111
Greaves, J. S., Holland, W. S., Moriarty-Schieven, G., et al.
1998, ApJL, 506, L133
Heppenheimer, T. A. 1978, A&A, 65, 421
Hillenbrand, L. A., Carpenter, J. M., Kim, J. S., et al. 2008,
Apj, 677, 630
Kalas, P., Graham, J. R., Chiang, E., et al. 2008, Science,
322, 1345
Kenyon, S. J. & Bromley, B. C. 2004, Aj, 127, 513
|
1611.09236 | 1 | 1611 | 2016-11-28T16:50:58 | An Overabundance of Low-density Neptune-like Planets | [
"astro-ph.EP"
] | We present a uniform analysis of the atmospheric escape rate of Neptune-like planets with estimated radius and mass (restricted to $M_{\rm p}<30\,M_{\oplus}$). For each planet we compute the restricted Jeans escape parameter, $\Lambda$, for a hydrogen atom evaluated at the planetary mass, radius, and equilibrium temperature. Values of $\Lambda\lesssim20$ suggest extremely high mass-loss rates. We identify 27 planets (out of 167) that are simultaneously consistent with hydrogen-dominated atmospheres and are expected to exhibit extreme mass-loss rates. We further estimate the mass-loss rates ($L_{\rm hy}$) of these planets with tailored atmospheric hydrodynamic models. We compare $L_{\rm hy}$ to the energy-limited (maximum-possible high-energy driven) mass-loss rates. We confirm that 25 planets (15\% of the sample) exhibit extremely high mass-loss rates ($L_{\rm hy}>0.1\,M_{\oplus}{\rm Gyr}^{-1}$), well in excess of the energy-limited mass-loss rates. This constitutes a contradiction, since the hydrogen envelopes cannot be retained given the high mass-loss rates. We hypothesize that these planets are not truly under such high mass-loss rates. Instead, either hydrodynamic models overestimate the mass-loss rates, transit-timing-variation measurements underestimate the planetary masses, optical transit observations overestimate the planetary radii (due to high-altitude clouds), or Neptunes have consistently higher albedos than Jupiter planets. We conclude that at least one of these established estimations/techniques is consistently producing biased values for Neptune planets. Such an important fraction of exoplanets with misinterpreted parameters can significantly bias our view of populations studies, like the observed mass--radius distribution of exoplanets for example. | astro-ph.EP | astro-ph |
MNRAS 000, 1–9 (2016)
Preprint 29 November 2016
Compiled using MNRAS LATEX style file v3.0
An overabundance of low-density Neptune-like planets
Patricio Cubillos1⋆, Nikolai V. Erkaev2, Ines Juvan1, Luca Fossati1, Colin P. Johnstone3,
Helmut Lammer1, Monika Lendl1, Petra Odert1, and Kristina G. Kislyakova1
1 Space Research Institute, Austrian Academy of Sciences, Schmiedlstrasse 6, A-8042 Graz, Austria.
2 Federal Research Center "Krasnoyarsk Science Center" SB RAS, "Institute of Computational Modelling".
3 University of Vienna, Department of Astrophysics, Turkenschanzstrasse 17, 1180 Vienna, Austria
Accepted XXX. Received YYY; in original form ZZZ
ABSTRACT
We present a uniform analysis of the atmospheric escape rate of Neptune-like plan-
ets with estimated radius and mass (restricted to Mp < 30 M⊕). For each planet
we compute the restricted Jeans escape parameter, Λ, for a hydrogen atom evalu-
ated at the planetary mass, radius, and equilibrium temperature. Values of Λ . 20
suggest extremely high mass-loss rates. We identify 27 planets (out of 167) that are
simultaneously consistent with hydrogen-dominated atmospheres and are expected
to exhibit extreme mass-loss rates. We further estimate the mass-loss rates (Lhy)
of these planets with tailored atmospheric hydrodynamic models. We compare Lhy
to the energy-limited (maximum-possible high-energy driven) mass-loss rates. We
confirm that 25 planets (15% of the sample) exhibit extremely high mass-loss rates
(Lhy > 0.1 M⊕Gyr−1), well in excess of the energy-limited mass-loss rates. This con-
stitutes a contradiction, since the hydrogen envelopes cannot be retained given the
high mass-loss rates. We hypothesize that these planets are not truly under such high
mass-loss rates. Instead, either hydrodynamic models overestimate the mass-loss rates,
transit-timing-variation measurements underestimate the planetary masses, optical
transit observations overestimate the planetary radii (due to high-altitude clouds), or
Neptunes have consistently higher albedos than Jupiter planets. We conclude that at
least one of these established estimations/techniques is consistently producing biased
values for Neptune planets. Such an important fraction of exoplanets with misin-
terpreted parameters can significantly bias our view of populations studies, like the
observed mass–radius distribution of exoplanets for example.
Key words: planets and satellites: atmospheres – planets and satellites: fundamental
parameters – hydrodynamics
1 INTRODUCTION
The Kepler Space Telescope mission has enabled the first
estimations of the abundance and size distribution of ex-
trasolar planets in our galaxy (e.g., Fressin et al. 2013;
Dressing & Charbonneau 2015). A large fraction of these
worlds have sizes intermediate between that of Earth and
Neptune, having no analog in our solar system. Further
follow-up studies have yielded mass estimates for a large
sample of Neptune-like planets (hereafter, considered as
those planets with Mp < 30M⊕ ≈ 2MNep), allowing us to
study the physical properties of exoplanets in a statistically
robust manner.
The planetary radius is inferred from optical tran-
typically corresponding to
light-curve observations,
sit
⋆ E-mail: [email protected]
c(cid:13) 2016 The Authors
level
for
a
clear
the ∼ 20 − 100 mbar
atmo-
sphere (e.g., Lopez & Fortney 2014; Lammer et al. 2016;
Lecavelier des Etangs et al. 2008). The planetary mass, in
turn, is inferred either from measurements of the stellar ra-
dial velocity (RV) or from transit timing variations (TTV)
due to interplanetary perturbations. By combining the mass
and radius constraints, we can infer bulk densities and com-
positions. For a given mass, the planetary radius does not
significantly depend on the interior composition, since it is
made of very incompressible materials. On the contrary, for
a given mass, the hydrogen envelope fraction strongly cor-
relates with the planetary radius (Lopez & Fortney 2014).
The mass–size distribution shows that a considerable
fraction of Neptune-like planets have significant envelopes
(a few percent of the core mass). Noteworthy, a couple
dozen of sub Neptunes show bulk densities far lower than
that of the solar-system giant planets. While it is plausi-
2
P. Cubillos et al.
ble that solid cores of a few M⊕ can accrete significant
gas envelopes before the protoplanetary disk dissipation
(Lee et al. 2014; Inamdar & Schlichting 2015; Stokl et al.
2016; Ginzburg et al. 2016), several studies find that after
the disk dispersal these gas envelopes endure significant mass
loss due to different mechanisms.
Cooling and photoevaporation, driven by the high-
energy stellar irradiation (XUV), produce moderate mass-
loss rates and atmospheric contraction. These mechanisms
act over timescales larger than ∼ 108 yr (e.g., Lopez et al.
2012; Owen & Wu 2013; Ginzburg et al. 2016). Hydrody-
namic and photochemical processes determine the composi-
tion of the upper atmosphere of these planets, and therefore
of the escaping particles Koskinen et al. (e.g., 2013a,b)
Using hydrodynamic simulations, Owen & Wu (2016)
found that during the first ∼Myr after the disk dispersal,
low-density planets exhibit extremely high thermally driven
mass-loss rates, dubbed "boil-off" (see also Lammer et al.
2016; Fossati et al. 2016). This mechanism consists of hy-
drodynamic thermal evaporation (Parker wind), driven by
the internal heat from the planet, and fueled by the stellar
continuum irradiation. During this regime, the atmosphere
quickly cools and contracts as it releases energy through the
escaped particles. In other words, the thermal energy ex-
ceeds the gravitational energy in atmospheric layers where
the density is high enough to lead to high mass-loss rates.
Consequently, the mass-loss rate exponentially decays until
the XUV-driven photoevaporation becomes the dominant
mass-loss mechanism (Stokl et al. 2015).
The extremely high mass-loss rates of this regime are
unsustainable for the atmospheres of Neptune-like planets
over Gyr timescales. Therefore, only young systems are ex-
pected to show thermally driven mass-loss rates in excess of
the XUV-driven mass-loss rates (Lammer et al. 2016).
Ultraviolet
transit
observations provide
Fossati et al.
loss on exoplanets
Linsky et al.
2010;
evidence
(e.g., Vidal-Madjar et al.
of mass
2003;
2010;
Ben-Jaffel & Ballester 2013). For example, Ly-α transit
observations (e.g., Vidal-Madjar et al. 2003; Kulow et al.
2014; Ehrenreich et al. 2015; Bourrier et al. 2016) show
transit depths much larger than in the optical. These ob-
servations are interpreted as absorption from an extended
region of escaping gas beyond the Roche lobe of the planet.
However, the translation from observations into mass-
loss rates is very model-dependent. Ly-α observations
show large offsets between the absorption features and
the center of the line (±100 km s−1), requiring addi-
tional assumptions to explain the observations. Among
the proposed mechanisms there are: natural broadening
from large-scale confinement of material
into a higher-
density region (Stone & Proga 2009; Owen & Adams 2014),
a combination of stellar radiation pressure and wind in-
teractions, where particles accelerate by charge exchange
in the wind-wind interaction region (Holmstrom et al.
2008; Tremblin & Chiang 2013; Kislyakova et al. 2014;
Christie et al. 2016), or inhomogeneities of the stellar disc
at Ly-α light (Llama et al. 2013).
Chamberlain & Hunten 1987). They generalized the Jeans
escape parameter for a hydrodynamic atmosphere subjected
to the gravitational perturbation from the host star. They
studied the upper atmosphere, between the 100 mbar level
and the Roche-lobe radius, using hydrodynamic models. By
deriving the generalized Jeans escape parameter across the
atmospheric layers, they were able to determine how stable
a planetary atmosphere is against evaporation.
Fossati et al. (2016) further defined the restricted Jeans
escape parameter:
Λ ≡
GMpmH
kBTeqRp
,
(1)
the Jeans escape parameter for a hydrogen atom evaluated
at the planetary mass (Mp), radius (Rp), and equilibrium
temperature (Teq), where mH is the mass of the hydrogen
atom, G is the gravitational constant, and kB is the Boltz-
mann constant. For hydrogen-dominated atmospheres, Λ is
an easy-to-calculate parameter that allows one to estimate
when the hydrodynamic mass-loss rate exceeds the XUV-
driven photoevaporation.
We note that, whereas the classical Jeans escape param-
eter is a variable as function of altitude in an atmosphere,
our definition of Λ is a particular value of it that works
as a global parameter for a given planet (no altitude de-
pendence). This is similar to the definition of Guillot et al.
(1996), but evaluated at the planetary equilibrium tempera-
ture. Using this specific definition, we empirically found the
boil-off threshold at Λ ∼ 20 (Fossati et al. 2016), equivalent
to the threshold of R = 0.1RBondi of Owen & Wu (2016).
Fossati et al. (2016) determined that planets with Λ .
20–40 should be experiencing extreme mass-loss rates by
comparing the hydrodynamic escape rate, Lhy, to the max-
imum XUV-driven escape rate, Len (estimated through the
energy-limited formula, Watson et al. 1981), over a range of
Λ scenarios with varying stellar type, planetary mass, and
planetary radius.
In this paper we follow up the studies of atmospheric
escape, modeling and estimating the present-day mass-loss
rates for a sample of known exoplanets. We compile a list of
the exoplanets with estimated masses and radii, and calcu-
late their restricted Jeans escape parameter, Λ (Section 2).
Then, we compute hydrodynamic models tailored for planets
suspected of being in the boil-off regime (Section 3). Finally
we discuss the observational and physical implications of our
findings (Section 4).
2 SAMPLE OF KNOWN NEPTUNES
We compiled our sample by collecting and crosschecking the
lists of exoplanets from the Nasa Exoplanet Archive1 the Ex-
oplanets Data Explorer2 (Han et al. 2014), and The Extra-
solar Planets Encyclopaedia3 (Schneider et al. 2012). We se-
lected the targets with measured planetary radii and masses,
whose mass is less than ∼ 2 Neptune masses (Mp < 30 M⊕).
We adopted stellar rotational angular velocity (Ωrot) from
1.1 Restricted Jeans Escape Parameter
Fossati et al.
(2016) described the thermal escape in
terms of the classical Jeans escape parameter (see e.g.,
1 http://exoplanetarchive.ipac.caltech.edu/
2 http://exoplanets.org/
3 http://exoplanet.eu/
MNRAS 000, 1–9 (2016)
McQuillan et al. (2013) and ages from Morton et al. (2016).
Our final sample consists of 167 planets (Table A1).
Fig. 1 shows the mass–radius–T eq distribution for the
planets in this sample. We calculated the equilibrium tem-
perature assuming zero Bond albedo and efficient day–night
energy redistribution. We split this figure into temperature
bins to show representative Λ contours for the planets.
This is a heterogeneous sample of planets. The large
majority of these planets (90% of the sample) were discov-
ered using the transit method from the Kepler Space Tele-
scope. The masses of these planets were estimated using the
TTV and RV methods (70% and 30% of the sample, respec-
tively), with only a few systems having both TTV and RV
constraints: Kepler-18, Kepler-89, K2-19, and WASP-47.
The distribution of planets in Fig. 1 reflects the selec-
tion biases from each observing method. Most RV planets
fall into the two panels with higher T eq because the RV
method favors planets orbiting closer to their host stars.
Furthermore, for a given planet size, the RV method tends
to find planets with higher mass while the sensitivity of
TTVs is more uniform (Weiss & Marcy 2014; Steffen 2016).
Jontof-Hutter et al. (2014) argued that larger planets have
deeper transits yielding more precise transit times.
Table A1 lists the observed and derived parameters for
each planet. We selected the planets consistent with an enve-
lope mass fraction larger than 1% (88 planets) by comparing
the observed radii to the models of Lopez & Fortney (2014),
for the given planetary masses, stellar ages, and incident
fluxes. From this sub sample, we identify 28 planets (17% of
the sample) with Λ < 20, and thus should exhibit extreme
mass-loss rates. We note that all of these systems are old (>1
Gyr). We observe some exceptional cases in system param-
eters, extremely low-bulk-density planets (ρp < 0.2 g cm−3)
like Kepler-33, Kepler-51, Kepler-79, Kepler-87, or Kepler-
177; and extremely unstable atmospheres (Λ < 5) like
Kepler-33 c or Kepler-51 b.
3 HYDRODYNAMIC RUNS
In this section we present hydrodynamical models of the
upper atmosphere of selected Neptune-like planets, which
allow us to directly estimate the present-day hydrodynamic
mass-loss rate. We further determine whether their mass-
loss rate is higher than the upper-limit XUV-driven mass-
loss rate. We covered a range of Λ from 0 to 50, including all
hydrogen-dominated atmospheres with Λ < 20, some higher-
density planets, and some planets with larger Λ.
To compute Lhy, we applied the one-dimensional upper-
atmosphere hydrodynamic model of Erkaev et al. (2016).
This model solves the hydrodynamic system of equations
for mass, momentum, and energy conservation, considering
the absorption of stellar XUV flux and accounting for the
particles' ionization, dissociation, recombination, and Ly-α
cooling4.
For the XUV absorption we assume a spherically sym-
metric distribution of density, deviations from this symme-
try do not seem to be essential. One could expect that differ-
ence between 1D and 3D models would be more pronounced
4 See the appendix in Erkaev et al. (2016) for a detailed descrip-
tion.
MNRAS 000, 1–9 (2016)
An overabundance of low-density Neptunes
3
for cases of larger Jeans escape parameter, when the hy-
drodynamic flow is driven mainly by the XUV flux, which
may have strong asymmetries. But in such case the mass-
loss rate is rather close to the energy-limited escape value,
as observed in existing 2D (Khodachenko et al. 2015) and
3D (Tripathi et al. 2015) simulations.
For hot Jupiters, our model produces similar mass-loss
rates as most other hydrodynamic models (Erkaev et al.
2016). For the Neptune-like planet GJ 436 b, our model pre-
dicts a mass-loss rate of 2×10−9 g s−1, consistent with the
values from Ehrenreich et al. (2015).
As in Tu et al. (2015) and Johnstone et al. (2015), we
calculate the stellar X-ray and EUV luminosities using the
scaling laws of Wright et al. (2011) to convert the stel-
lar rotation rates and masses into X-ray luminosities, and
Sanz-Forcada et al. (2011) to convert the X-ray luminosi-
ties into EUV luminosities. When the stellar rotation rate
is not known, we use the gyrochronological relation of
Mamajek & Hillenbrand (2008) to convert stellar age into
rotation rate.
We calculate the upper-limit XUV-driven escape
using the energy-limited formula (Watson et al. 1981;
Erkaev et al. 2007):
Len =
πηRp(Reff
XUV)2FXUV
GMpmHK
,
(2)
with η = 15% the net heating efficiency (Shematovich et al.
2014), Reff
XUV the radius where the bulk of the XUV energy is
absorbed, and K the potential energy reduction factor due
to stellar tides, which depends on the Roche-lobe boundary
radius.
It is important to remark that both the stellar flux
and the planetary parameters vary over their respective
lifespan,
inducing a variation in the planetary mass-loss
rate with time (see, e.g., Lecavelier des Etangs et al. 2004;
Lecavelier Des Etangs 2007). In particular, for the boil-off
regime, Owen & Wu (2016) argue that the mass-loss rate is
exponentially sensitive to the planetary radius at small Λ.
Thus, during boil-off, strong cooling and contraction quickly
reduce Rp (and hence Lhy) in timescales of ∼ 1 Myr until
Λ ∼ 20. The XUV stellar fluxes we compute take into ac-
count the age of the star, whereas the planetary radii and
masses are taken directly from the reported values. There-
fore, the mass-loss rates we compute should be considered
as a present-day snapshot of their values.
Table A1 and Fig. 2 show the hydrodynamic and
energy-limited mass-loss rates for each modeled planet. Note
that the estimated Lhy is the total mass-loss rate, including
photoevaporation. Thus, we can infer whether photoevapo-
ration (Lhy ≈ Len) or boil-off (Lhy ≫ Len) is the dominant
mass-loss mechanism. Given the variability of Lhy and Len
with the adopted model parameters (FXUV and η), we con-
sider that planets with Lhy/Len > 2.5 are in the boil-off
regime (Lammer et al. 2016).
We confirm that Λ is a good indicator of extreme mass-
loss rates (Fig. 2, top panel). For Λ < 15, Lhy/Len is strongly
correlated with Λ, exponentially increasing with decreasing
Λ. In the intermediate range 15 < Λ < 25, there is a tran-
sition from boil off to photoevaporation-dominated mass-
loss rate. Then, for Λ > 25, Lhy/Len remains constant as
the mass-loss rate is dominated by the XUV-driven pho-
toevaporation. We find no significant correlation of Λ with
4
P. Cubillos et al.
0
.
5
5
.
1
10
)
⊕
R
(
i
s
u
d
a
R
8
6
4
2
0
10
5
.
1
)
⊕
R
(
i
s
u
d
a
R
8
6
4
2
0
10
)
⊕
R
(
i
s
u
d
a
R
8
6
4
5
.
1
2
0
0
0.0
1
20.0
30.0
300 K < T < 550 K
0
.
5
0.0
1
550 K < T < 650 K
3 0.0
20.0
0.0
5
0 . 7
0 . 0
8
1 . 6
5
.
1
5.5
1 5 0 . 0
0 . 0
5
0 . 7
1 . 6
8 0 . 0
5.5
1 5 0 . 0
20.0
650 K < T < 800 K
20.0
800 K < T < 950 K
0
.
5
0.0
1
0
.
5
0.0
1
3 0.0
0 . 0
0 . 7
5
1 . 6
8 0 . 0
5.5
1 5 0 . 0
5
.
1
0 . 0
3
0 . 7
0 . 0
5
1 . 6
5.5
8 0 . 0
1 5 0 . 0
10.0
950 K < T < 1100 K
2 0.0
5.0
5
.
1
5.0
10.0
1100 K < T < 2300 K
0 . 0
3
5.5
0 . 7
1 . 6
5 0 . 0
8 0 . 0
1 5 0 . 0
0 . 0
2
0 . 7
1 . 6
0 . 0
3
150.0
5.5
5 0 . 0
8 0 . 0
5
10
15
20
25
30
0
5
10
15
20
25
30
Mass (M⊕)
Mass (M⊕)
Figure 1. Mass–radius–T eq distribution of the known Neptune-like planets. The blue and green markers denote RV- and TTV-measured
masses, respectively. From top to bottom, the brown solid lines denote constant-density curves corresponding to Saturn's, Neptune's, and
Earth's bulk density in g cm−3, respectively. The background color contour denotes the value of Λ for the mean equilibrium temperature
of the planets in each panel: 455, 600, 715, 870, 1025, and 1440 K. Note that the Λ-contours work only as a guideline, Table A1 gives
the specific values for each planet.
MNRAS 000, 1–9 (2016)
102
2e6
n
e
L
/
y
h
L
101
100
10-1
0
104
2e7
102
)
1
−
r
y
G
⊕
M
(
100
y
h
L
10-2
10
20
30
40
50
60
20.0
17.5
15.0
12.5
10.0
7.5
5.0
2.5
)
⊕
M
(
s
s
a
M
0
10
20
40
50
60
30
Λ
Figure 2. Top: Ratio between the hydrodynamic and energy-
limited mass-loss rate vs. Λ. The color scale denotes the planetary
mass. The background color marks the regions with dominant ex-
treme thermal mass-loss rate (gray) and XUV-driven photoevap-
oration (white). The labeled triangle denotes the mass-loss values
for the extreme case of Kepler-33 c. Bottom: Hydrodynamic mass-
loss rate vs. Λ.
other observed parameters. Furthermore, 19 of the 27 pre-
viously identified planets with hydrogen-dominated atmo-
spheres and Λ < 20 satisfy Lhy/Len > 2.5. Six additional
low-density planets with Λ ≤ 25 also present an excess mass-
loss rate. The bottom panel of Fig. 2 shows the computed
hydrodynamic mass-loss rates. All planets with excess mass-
loss rates (Lhy/Len > 2.5) show hydrodynamic mass-loss
rates above 0.1 M⊕Gyr−1. At this rate, these planets should
have already lost their hydrogen atmospheres.
The scatter in Lhy (Fig. 2, bottom panel) seems to be
mainly correlated with the planetary mass: at a given Λ
more massive planets have higher mass-loss rates. On the
contrary, the main source for the scatter in Lhy/Len (Fig.
2, top panel) is unclear; these variations seem to be a more
complex combination of several factors, e.g., Roche-lobe ra-
dius, incident stellar flux, XUV absorption radius, etc.
4 DISCUSSION
The observed low bulk densities of many Neptune-like plan-
ets require a significant hydrogen envelope fraction. How-
ever, their estimated restricted Jeans escape parameter, Λ,
and further hydrodynamic modeling indicate that 25 plan-
ets in our sample (15%) should be experiencing extremely
high mass-loss rates (Lhy > 0.1 M⊕Gyr−1) in excess of the
maximum XUV-driven photoevaporation (Lhy/Len > 2.5).
Considering the age of the systems (& Gyr), it is im-
probable that these Neptunes have retained their hydro-
gen envelopes until the present day. In addition, the short
timescale of the boil-off regime (. Myr) makes it unlikely
that we are observing a transient phenomenon by chance.
This contradiction leads us to consider biases on the mass-
loss model estimations, the measured physical parameters,
or the physical interpretation of the observations.
MNRAS 000, 1–9 (2016)
An overabundance of low-density Neptunes
5
4.1 Hydrodynamic-model Bias?
Validating model mass-loss rates is challenging because to
date there are no direct observational measurements. UV
transit observations depend as much on the hydrodynamic
models inside the planetary Roche lobe as in the particle
dynamics after. This would require a treatment of a number
of physical processes that are largely unconstrained (e.g.,
magnetic fields, stellar radiation pressure, wind-wind inter-
action).
The heating efficiency rate η is one of the least con-
strained parameters in our model, for which we assume a
value of 0.15. A detailed calculation of η is complicated, as it
varies with altitude and it requires a kinetic approach consid-
ering many chemical reactions. Shematovich et al. (2014) es-
timated values of η between 0.1 and 0.2 for hot Jupiters. As-
suming larger heating efficiency one can increase the mass-
loss rates; however, Owen & Jackson (2012) discarded values
larger than η = 0.4. Salz et al. (2016) found little variation
of the heating efficiency (η ≈ 0.1–0.3) with gravitational
potential, except for the most compact and massive planets.
One process that we do not consider is magnetic fields.
Owen & Adams
(2015)
found that magnetic fields can suppress mass-loss rates by
approximately an order of magnitude by confining the out-
flowing material
into smaller opening angles around the
poles.
(2014) and Khodachenko et al.
4.2 Planetary-mass Bias?
One explanation to this contradiction is that TTV analy-
ses are underestimating the planetary masses, a possibility
already considered by Weiss & Marcy (2014). A more mas-
sive planet would have a stronger gravitational pull, decreas-
ing the mass-loss rate. Steffen (2016) argues that the mass
measurements derived from RV and TTV are comparably
reliable since the physics behind TTV measurements (grav-
ity) is well understood. However, given the complexity and
rapid-pace development of exoplanet RV and TTV analy-
ses, as data-reduction techniques improve, their mass esti-
mations continuously get overturned or refined. For example,
TTV-estimated masses for Kepler-114 c (Xie 2014), Kepler-
231 c (Kipping et al. 2014), and Kepler-56 b (Huber et al.
2013) differ by a factor of ten from a nearly contempora-
neous estimation (Hadden & Lithwick 2014). Cases of ex-
tremely high densities (e.g., Kepler-327 c, ρp ≈ 70 g cm−3
Hadden & Lithwick 2014) or even negative RV-estimated
masses (e.g., Marcy et al. 2014) serve as reminders to use
the data-reduction and statistical tools with caution. One
should incorporate all available prior information (e.g., in a
Bayesian posterior sampling) to avoid results at odds with
physical principles.
In our sample, with the exception of Kepler-94 b, all
boil-off planets have TTV-determined masses. Ideally, com-
paring the TTV and RV results for a given system would
test the reliability of these techniques. However, due to the
observing limitations of each method, only a couple of sys-
tems have allowed for independent TTV and RV analyses.
For example, for the Kepler-89 system, Masuda et al. (2013)
found discrepancies between the TTV and RV-mass esti-
mations. On the contrary, for the Kepler-18 and WASP-
6
P. Cubillos et al.
47 systems, TTV and RV analyses return consistent results
(Cochran et al. 2011; Dai et al. 2015).
If we adopted masses such that Λ ≡ 20 (M20, keeping all
other parameters fixed), 13 of the 19 low-density extreme-
mass-loss planets with Λ < 20.0 would have a mass cor-
rection larger than 1σM, with a mean of 5.0σM (see Table
1).
4.3 Planetary-radius Bias?
Another possibility is that we are misinterpreting the ob-
served transit radius. For clear atmospheres in Neptune-
like planets, the observed optical transit radius corre-
sponds to the ∼20–100 mbar level (Lopez & Fortney 2014;
Lammer et al. 2016). However, if a planet has high-altitude,
optically thick clouds/hazes, the transit radius would be
overestimating the planetary radius. If this is the case,
then the true 100 mbar altitude would correspond to a
smaller radius than the observed radius, yielding moder-
ate mass-loss rates. Spectroscopic analyses already indicate
that many Neptune and sub-Neptune sized planets show
flat transmission spectra, consistent with gray opacity hazes
or clouds (e.g., Kreidberg et al. 2014; Knutson et al. 2014;
Ehrenreich et al. 2014).
If we adopted radii such that Λ ≡ 20 (R20, keeping
all other parameters fixed), 18 of the 19 low-density boil-
off planets with Λ < 20.0 would have a radius correc-
tion larger than 1σR, with a mean of 8.5σR (see Table 1).
Adopting R20 as the 100 mbar pressure level, we computed
hydrostatic-equilibrium pressure–radius profiles to calculate
the observed transit photospheric pressure, pphot = p(Rp).
Table 1 shows that a third of the planets could have photo-
spheres at pressures between 10−3 bar and 10−7 bar, com-
patible with the location of cloud decks on exoplanets.
When we plot the planetary surface gravity, log g, vs.
equilibrium temperature for the planets in our sample (Fig.
3, top panel), we see that all boil-off planets are grouped in
the log g < 3.0 region. While it is not surprising that these
planets are concentrated at low log g values (given the sim-
ilarity in mass and radius dependency of log g ∝ Mp R−2
and Λ ∝ Mp R−1
p ), it is interesting that this region limits
closely to the boundary beweeen strong and weak J-band
H2O features observed in hot Jupiters, associated to cloudy
vs. clear atmospheres, respectively (Stevenson 2016). Re-
gardless of whether there is a link between cloud processes
on hot Jupiters and Neptune-like planets -two completely
different samples-, this graph helps to identify planets with
a higher prominence of clouds/hazes. Note that the derived
surface gravity of these planets is also affected by the obser-
vational biases. Fig. 3 (bottom panel) shows the corrected
log g–T eq position of the boil-off planets if we assumed R20
as the 100 mbar radius.
p
An important remark to highlight is that, while our
analysis suggests that boil-off planets can have instead mod-
erate mass-loss rates if they have cloudy atmospheres, we
cannot claim that the remaining planets have clear atmo-
spheres. They can either host clear or cloudy atmospheres.
This occurs because our analysis is based on the domi-
nant effect of hydrogen in the observed planetary radius.
There can be cloudy planets with secondary atmospheres,
or even hydrogen-dominated atmospheres hosting lower-
Unmodeled
Lhy/Len <2.0
Lhy/Len >2.0
Hot-super-Earth desert
2000
)
K
(
q
e
T
1500
1000
500
1.5
2.0
2.5
3.0
3.5
4.0
)
K
(
q
e
T
1200
1050
900
750
600
450
300
1.5
2.0
2.5
3.0
log g (cm s−2 )
3.5
4.0
Figure 3. Planetary surface gravity vs. equilibrium-temperature.
The orange, white, and blue circles (top panel) denote the extreme
mass-loss rate, moderate mass-loss rate, and unmodeled planets,
respectively. The gray and white background colors denote the
region where planets have clear and cloudy-atmospheres, respec-
tively, derived by Stevenson (2016) for hot Jupiters. The horizon-
tal dashed line denotes the lowest incident-stellar-flux boundary
(650 F⊕) of the hot-super-Earth desert (Lundkvist et al. 2016).
The bottom panel shows the expected log g of the boil-off plan-
ets if their radii were corrected to have Λ ≡ 20.0 (brown circles)
relative to their initial position (orange circles).
altitude clouds that do not alter much the observed transit
radius.
4.4 Bond-albedo Bias?
The cloudy-atmospheres scenario has a second consequence
on the properties of these planets. The high-altitude
cloud/haze layer would reflect a larger fraction of the in-
cident stellar irradiation, decreasing the amount of energy
deposited in the atmosphere, and consequently decreasing
the estimated mass-loss rates. Low-density sub Neptunes (on
average) would then differ from hot Jupiters. Studies of the
aggregate sample of hot Jupiters show low geometric opti-
cal albedos (Ag ≈ 0.1) and somewhat higher Bond albedos
(AB = 0.4 ± 0.1, Schwartz & Cowan 2015), suggesting that
hot Jupiters have a layer of optical absorbers above an infra-
red reflective cloud deck. Our cloudy-atmosphere case aligns
well with the higher geometric albedos (Ag ≈ 0.16 − 0.3)
found on super-Earth planets (Rp < 2.5R⊕, Demory 2014).
Table 1 shows the required values of the Bond albedo
such that Λ ≡ 20.0 (A20, keeping all other parameters fixed).
Most of the planets in this list would require Bond albedos
higher than any observed value from solar or extrasolar gas-
giant planet.
If we confirm the cloudy nature of
these atmo-
MNRAS 000, 1–9 (2016)
Table 1. Low-density boil-off planet parameters. Λ, Mp, Rp, and T eq are as in Table A1. M20, R20, and A20 are the modified mass,
radius, and bond albedo such that Λ = 20. pphot is the photospheric pressure corresponding to Rp when we adopt R20 as the 100 mbar
pressure level.
An overabundance of low-density Neptunes
7
Planet name
Λ
Kepler-11 c
Kepler-11 f
Kepler-114 d
Kepler-177 b
Kepler-177 c
Kepler-254 c
Kepler-29 b
Kepler-29 c
Kepler-305 c
Kepler-307 b
Kepler-307 c
Kepler-33 c
Kepler-33 d
Kepler-51 b
Kepler-51 c
Kepler-51 d
Kepler-79 d
Kepler-79 e
Kepler-87 c
8.8
10.3
18.1
7.6
13.7
13.7
11.7
12.0
17.2
8.5
5.2
1.7
7.1
4.2
7.7
15.1
10.1
16.4
17.7
Mp
M⊕
2.86+2.86
−1.59
1.91±0.95
3.81±1.59
1.91+0.64
−0.32
7.63+3.50
−3.18
3.20+3.70
−2.70
4.50±1.50
4.00±1.30
6.04+2.54
−2.22
3.18±0.64
1.59+0.64
−0.32
0.80+2.50
−0.70
4.70±2.00
2.22+1.59
−0.95
4.13±0.32
7.63±0.95
6.04+2.10
−1.59
4.13+1.21
−1.11
6.40±0.80
M20
M⊕
6.5
3.7
4.2
5.0
11.1
4.7
7.7
6.6
7.0
7.4
6.1
9.4
13.3
10.5
10.8
10.1
12.0
5.0
7.2
Rp
R⊕
2.87+0.04
−0.06
2.49+0.04
−0.07
2.54±0.28
2.91+1.53
−0.30
7.10+3.71
−0.72
2.09+0.84
−0.16
3.35±0.22
3.14±0.20
3.30+0.82
−0.33
3.20+1.20
−0.46
2.81+1.05
−0.42
3.20±0.30
5.35±0.49
7.10±0.30
9.01+2.81
−1.71
9.71±0.50
7.17+0.13
−0.16
3.49±0.13
6.15±0.09
R20
R⊕
1.26
1.29
2.30
1.10
4.87
1.44
1.95
1.89
2.83
1.36
0.74
0.27
1.89
1.50
3.45
7.34
3.61
2.87
5.46
pphot
bar
6.8 × 10−13
2.4 × 10−11
1.4 × 10−03
4.7 × 10−14
5.5 × 10−08
5.5 × 10−08
5.0 × 10−10
1.2 × 10−09
1.4 × 10−04
4.1 × 10−13
2.3 × 10−16
6.7 × 10−20
1.3 × 10−14
2.3 × 10−17
5.3 × 10−14
1.3 × 10−06
1.4 × 10−11
2.7 × 10−05
5.1 × 10−04
T eq
K
859
562
628
655
594
845
872
802
808
883
818
1113
941
561
453
394
633
546
444
A20
0.96
0.93
0.32
0.98
0.78
0.78
0.88
0.87
0.46
0.97
0.99
0.99
0.98
0.99
0.98
0.67
0.94
0.54
0.38
spheres, this analysis places strong constraints on the at-
mospheric properties of the ensemble of low-density sub
Neptunes. These planets would present high-altitude opaque
clouds/hazes that increase the observed radii. Observation-
ally, the value of their restricted Jeans escape parameter Λ,
and also their position in the log g–T eq plane, would be sim-
ple and direct indicators of the cloud/haze prominence on
the atmospheres of these planets.
On a different but still
interesting topic, the dis-
tribution of planets in the log g–T eq plane (Fig. 3 top
panel) agrees well with the proposed hot-super-Earth desert
(Lundkvist et al. 2016). The boundary at 650 times the in-
cident flux on Earth (F⊕) directly translates into an equilib-
rium temperature of ∼ 1400 K. From the planets with known
mass and radius, we observe a dearth of planets above this
line. Unlike the Rp–F⊕ figure from Lundkvist et al. (2016),
this figure further constrains the parameter space of the
planetary physical properties of the desert by adding infor-
mation from the masses (through the surface gravity).
Furthermore, by considering the boil-off mass-loss
regime, our results (Fig. 3 bottom panel) suggest that the
exoplanet desert extends even further. Be it the mass or the
radius of the planets being misinterpreted, boil-off planets
should have a higher surface gravity, further depleting a re-
gion below the Teq = 1400 K boundary, down to ∼ 600 K.
5 CONCLUSIONS
We studied the mass-loss rates of the sample of 167 con-
firmed planets with masses Mp < 30M⊕. For 15% of
the planets in this sample, we found that their planetary
MNRAS 000, 1–9 (2016)
mass, radius, and equilibrium temperature lead to para-
doxical results. Their observed low bulk densities imply
hydrogen-dominated atmospheres. However, hydrodynami-
cal models indicate extremely high mass-loss rates (Lhy >
0.1 M⊕Gyr−1), thus being unable to retain their hydrogen
envelopes (boil-off regime).
Our hydrodynamical models' mass-loss rates agree well
with state-of-the-art models from the literature. Unfortu-
nately, current observations do not offer a direct, model-
independent validation of escape rates. Magnetic fields (how-
ever largely unknown) could offer a solution of this paradox
by decreasing the predicted mass-loss rates (Owen & Adams
2014; Khodachenko et al. 2015). Alternatively, the boil-off
timescales of ∼ 106 yr (Owen & Wu 2016) could have been
grossly underestimated. We need more detailed studies of
the mass-loss time evolution during the boil-off regime.
If the model mass-loss estimations are correct, we hy-
pothesize that either the TTV studies are underestimat-
ing the planetary masses or that these planets have high-
altitude clouds/hazes producing overestimated transit radii,
high albedos, or both. The cloudy/hazy scenario is further
supported by the higher albedo found for super Earths,
and a similar correlation as that between log g–T eq and
clear/cloudy atmospheres for hot Jupiters.
Whether we are misinterpreting their observed tran-
sit radii or underestimating their TTV masses, a fraction
of ∼10–20% Neptunes with biased parameters will have a
large impact on the population studies of the mass–radius
relationship. Both empirical laws and physically motivated
models tracing the observed mass-radius distribution of Nep-
tunes would be disrupted by this sub-sample, simply because
the observed values are not representative of the underlying
8
P. Cubillos et al.
properties of exoplanets. This could explain, for example,
why studies like Wolfgang & Lopez (2015) do not find a di-
rect mass–radius relationship for sub Neptune planets. In the
future, population studies should consider the added com-
plexity introduced by this rather large sub-sample of planets.
Conclusively solving this puzzle involves a multitude
of studies, from the confirmation of low-densities with RV
measurements, to the consistent modeling of high-altitude
clouds/hazes. This study generates many open questions for
the future, for example, can clouds/hazes form and remain at
such high altitudes and under strong stellar irradiation con-
ditions? Or how do reflective clouds affect the atmospheric
albedo, and hence mass loss?
ACKNOWLEDGMENTS
We thank contributors to SciPy, Matplotlib (Hunter 2007),
the Python Programming Language, and contributors to the
free and open-source community. We thank the anonymous
referee for comments that significantly improved the qual-
ity of the paper. This research has made use of the Ex-
oplanet Orbit Database and the Exoplanet Data Explorer
at exoplanets.org and the NASA Exoplanet Archive, which
is operated by the California Institute of Technology, un-
der contract with the National Aeronautics and Space Ad-
ministration under the Exoplanet Exploration Program. We
acknowledge the Austrian Forschungsforderungsgesellschaft
FFG projects "RASEN" P847963 and "TAPAS4CHEOPS"
P853993, the Austrian Science Fund (FWF) NFN projects
S11607-N16 and S11604-N16, and the FWF project P27256-
N27. N.V.E. acknowledges support by the RFBR grant No.
15-05-00879-a and No. 16-52-14006 ANF-a.
REFERENCES
Almenara J. M., et al., 2015, A&A, 581, L7
Alonso R., et al., 2014, A&A, 567, A112
Bakos G. ´A., et al., 2010, ApJ, 710, 1724
Barros S. C. C., et al., 2014, A&A, 569, A74
Barros S. C. C., et al., 2015, MNRAS, 454, 4267
Becker J. C., Vanderburg A., Adams F. C., Rappaport S. A.,
Schwengeler H. M., 2015, ApJ, 812, L18
Ben-Jaffel L., Ballester G. E., 2013, A&A, 553, A52
Berta-Thompson Z. K., et al., 2015, Nature, 527, 204
Biddle L. I., et al., 2014, MNRAS, 443, 1810
Borucki W. J., et al., 2010, ApJ, 713, L126
Bourrier V., Lecavelier des Etangs A., Ehrenreich D., Tanaka
Y. A., Vidotto A. A., 2016, A&A, 591, A121
Carter J. A., et al., 2012, Science, 337, 556
Chamberlain J. W., Hunten D. M., 1987, Theory of planetary
atmospheres. An introduction to their physics andchemistry.
Christie D., Arras P., Li Z.-Y., 2016, ApJ, 820, 3
Cochran W. D., et al., 2011, ApJS, 197, 7
Dai F., et al., 2015, ApJ, 813, L9
Demory B.-O., 2014, ApJ, 789, L20
Demory B.-O., et al., 2011, A&A, 533, A114
Dressing C. D., Charbonneau D., 2015, ApJ, 807, 45
Dressing C. D., et al., 2015, ApJ, 800, 135
Ehrenreich D., et al., 2014, A&A, 570, A89
Ehrenreich D., et al., 2015, Nature, 522, 459
Erkaev N. V., Kulikov Y. N., Lammer H., Selsis F., Langmayr D.,
Erkaev N. V., Lammer H., Odert P., Kislyakova K. G., John-
stone C. P., Gudel M., Khodachenko M. L., 2016, MNRAS,
460, 1300
Espinoza N., et al., 2016, ApJ, 830, 43
Fossati L., et al., 2010, ApJ, 714, L222
Fossati L., Erkaev N. V., Lammer H., Cubillos P., Juvan I.,
Kislyakova K. G., Lendl M., Odert P., 2016, A&A, Submitted.
Fressin F., et al., 2013, ApJ, 766, 81
Gautier III T. N., et al., 2012, ApJ, 749, 15
Gettel S., et al., 2016, ApJ, 816, 95
Gilliland R. L., et al., 2013, ApJ, 766, 40
Ginzburg S., Schlichting H. E., Sari R., 2016, ApJ, 825, 29
Guillot T., Burrows A., Hubbard W. B., Lunine J. I., Saumon D.,
1996, ApJ, 459, L35
Hadden S., Lithwick Y., 2014, ApJ, 787, 80
Hadden S., Lithwick Y., 2016, ApJ, 828, 44
Han E., Wang S. X., Wright J. T., Feng Y. K., Zhao M., Fakhouri
O., Brown J. I., Hancock C., 2014, PASP, 126, 827
Harpsøe K. B. W., et al., 2013, A&A, 549, A10
Hartman J. D., et al., 2011, ApJ, 728, 138
Holmstrom M., Ekenback A., Selsis F., Penz T., Lammer H., Wurz
P., 2008, Nature, 451, 970
Huber D., et al., 2013, Science, 342, 331
Hunter J. D., 2007, Computing In Science & Engineering, 9, 90
Inamdar N. K., Schlichting H. E., 2015, MNRAS, 448, 1751
Johnstone C. P., et al., 2015, ApJ, 815, L12
Jontof-Hutter D., Lissauer J. J., Rowe J. F., Fabrycky D. C.,
2014, ApJ, 785, 15
Jontof-Hutter D., Rowe J. F., Lissauer J. J., Fabrycky D. C., Ford
E. B., 2015, Nature, 522, 321
Jontof-Hutter D., et al., 2016, ApJ, 820, 39
Khodachenko M. L., Shaikhislamov I. F., Lammer H., Prokopov
P. A., 2015, ApJ, 813, 50
Kipping D. M., Nesvorn´y D., Buchhave L. A., Hartman J., Bakos
G. ´A., Schmitt A. R., 2014, ApJ, 784, 28
Kislyakova K. G., Holmstrom M., Lammer H., Odert P., Kho-
dachenko M. L., 2014, Science, 346, 981
Knutson H. A., Benneke B., Deming D., Homeier D., 2014,
Nature, 505, 66
Koskinen T. T., Harris M. J., Yelle R. V., Lavvas P., 2013a, Icarus,
226, 1678
Koskinen T. T., Yelle R. V., Harris M. J., Lavvas P., 2013b,
Icarus, 226, 1695
Kreidberg L., et al., 2014, Nature, 505, 69
Kulow J. R., France K., Linsky J., Loyd R. O. P., 2014, ApJ,
786, 132
Lammer H., et al., 2016, preprint, (arXiv:1605.03595)
Lecavelier Des Etangs A., 2007, A&A, 461, 1185
Lecavelier des Etangs A., Vidal-Madjar A., McConnell J. C.,
H´ebrard G., 2004, A&A, 418, L1
Lecavelier des Etangs A., Pont F., Vidal-Madjar A., Sing D., 2008,
A&A, 481, L83
Lee E. J., Chiang E., Ormel C. W., 2014, ApJ, 797, 95
Linsky J. L., Yang H., France K., Froning C. S., Green J. C.,
Stocke J. T., Osterman S. N., 2010, ApJ, 717, 1291
Lissauer J. J., et al., 2013, ApJ, 770, 131
Llama J., Vidotto A. A., Jardine M., Wood K., Fares R., Gombosi
T. I., 2013, MNRAS, 436, 2179
Lopez E. D., Fortney J. J., 2014, ApJ, 792, 1
Lopez E. D., Fortney J. J., Miller N., 2012, ApJ, 761, 59
Lundkvist M. S., et al., 2016, Nature Communications, 7, 11201
Maciejewski G., Niedzielski A., Nowak G., Pall´e E., Tingley B.,
Errmann R., Neuhauser R., 2014, Acta Astronomica, 64, 323
Mamajek E. E., Hillenbrand L. A., 2008, ApJ, 687, 1264
Marcy G. W., et al., 2014, ApJS, 210, 20
Masuda K., 2014, ApJ, 783, 53
Masuda K., Hirano T., Taruya A., Nagasawa M., Suto Y., 2013,
Jaritz G. F., Biernat H. K., 2007, A&A, 472, 329
ApJ, 778, 185
MNRAS 000, 1–9 (2016)
An overabundance of low-density Neptunes
9
McQuillan A., Mazeh T., Aigrain S., 2013, ApJ, 775, L11
Morton T. D., Bryson S. T., Coughlin J. L., Rowe J. F., Ravichan-
dran G., Petigura E. A., Haas M. R., Batalha N. M., 2016,
ApJ, 822, 86
Motalebi F., et al., 2015, A&A, 584, A72
Moutou C., et al., 2014, MNRAS, 444, 2783
Nesvorn´y D., Kipping D., Terrell D., Hartman J., Bakos G. ´A.,
Buchhave L. A., 2013, ApJ, 777, 3
Ofir A., Dreizler S., Zechmeister M., Husser T.-O., 2014, A&A,
561, A103
Owen J. E., Adams F. C., 2014, MNRAS, 444, 3761
Owen J. E., Jackson A. P., 2012, MNRAS, 425, 2931
Owen J. E., Wu Y., 2013, ApJ, 775, 105
Owen J. E., Wu Y., 2016, ApJ, 817, 107
Pepe F., et al., 2013, Nature, 503, 377
Petigura E. A., et al., 2016, ApJ, 818, 36
Salz M., Schneider P. C., Czesla S., Schmitt J. H. M. M., 2016,
A&A, 585, L2
Sanchis-Ojeda R., et al., 2012, Nature, 487, 449
Sanz-Forcada J., Micela G., Ribas I., Pollock A. M. T., Eiroa C.,
Velasco A., Solano E., Garc´ıa- ´Alvarez D., 2011, A&A, 532, A6
Schmitt J. R., et al., 2014, ApJ, 795, 167
Schneider J., Le Sidaner P., Savalle R., Zolotukhin I., 2012, in
Ballester P., Egret D., Lorente N. P. F., eds, Astronomical
Society of the Pacific Conference Series Vol. 461, Astronomical
Data Analysis Software and Systems XXI. p. 447
Schwartz J. C., Cowan N. B., 2015, MNRAS, 449, 4192
Shematovich V. I., Ionov D. E., Lammer H., 2014, A&A, 571, A94
Steffen J. H., 2016, MNRAS, 457, 4384
Stevenson K. B., 2016, ApJ, 817, L16
Stokl A., Dorfi E., Lammer H., 2015, A&A, 576, A87
Stokl A., Dorfi E. A., Johnstone C. P., Lammer H., 2016, ApJ,
825, 86
Stone J. M., Proga D., 2009, ApJ, 694, 205
Tremblin P., Chiang E., 2013, MNRAS, 428, 2565
Tripathi A., Kratter K. M., Murray-Clay R. A., Krumholz M. R.,
2015, ApJ, 808, 173
Tu L., Johnstone C. P., Gudel M., Lammer H., 2015, A&A,
577, L3
Van Grootel V., et al., 2014, ApJ, 786, 2
Vanderburg A., et al., 2015, ApJ, 800, 59
Vidal-Madjar A., Lecavelier des Etangs A., D´esert J.-M., Ballester
G. E., Ferlet R., H´ebrard G., Mayor M., 2003, Nature,
422, 143
Watson A. J., Donahue T. M., Walker J. C. G., 1981, Icarus,
48, 150
Weiss L. M., Marcy G. W., 2014, ApJ, 783, L6
Weiss L. M., et al., 2013, ApJ, 768, 14
Weiss L. M., et al., 2016, ApJ, 819, 83
Wolfgang A., Lopez E., 2015, ApJ, 806, 183
Wright N. J., Drake J. J., Mamajek E. E., Henry G. W., 2011,
ApJ, 743, 48
Xie J.-W., 2013, ApJS, 208, 22
Xie J.-W., 2014, ApJS, 210, 25
APPENDIX A: PARAMETERS OF THE
KNOWN NEPTUNE PLANETS
Table A1 list the observed and derived parameters for the
known sample of Neptune-like planets.
This paper has been typeset from a TEX/LATEX file prepared by
the author.
MNRAS 000, 1–9 (2016)
10
P. Cubillos et al.
Table A1. Observed and derived parameters for the sub-Neptune planet sample. This table only includes planets with estimated mass
(Mp) and transit radius (Rp), with Mp < 30M⊕. The equilibrium temperature (T eq) assumes zero Bond albedo and efficient day–night
energy redistribution. The restricted Jeans escape parameter (Λ) comes from Eq. (1). The planetary mean density (ρp) assumes the
observed mass and transit radius. FXUV (section 3) is the stellar XUV-flux received by the planet for the given orbital semi-major axis
(a), stellar mass (Ms), age, and rotational angular velocity (Ωrot). Lhy and Len (section 3) are the hydrodynamic and energy-limited
XUV-driven (Len) mass-loss rates of the selected hydrodynamic-modeled planets (mostly low-density planets with Λ < 20).
305 180.8
545 101.6
Λ
15.6
18.6
60.8
15.6
36.1
42.7
18.4
35.3
38.5
62.9
47.6
22.6
20.6
34.0
50.8
34.1
61.3
26.0
8.9
78.8
32.9
35.0
50.3
23.4
45.5
11.6
26.8
36.8
56.5
8.6
8.8
24.9
22.6
10.3
58.1
18.9
18.1
45.5
Name
55 Cnc e
CoRoT-22 b
CoRoT-24 c
Mp
M⊕
Rp
R⊕
Teq
K
2.08±0.16 1957
8.38±0.39
4.88+0.17
12.08+13.99
−0.39
−8.90
27.97±11.12 4.94±0.45
1007
706
CoRoT-7 b
5.72±0.95
1.58±0.07 1760
EPIC 2037 b 21.01±5.40
5.69±0.56
EPIC 2037 c
27.00±6.90
7.83±0.72
GJ 1132 b
GJ 1214 b
1.62±0.54
1.16±0.11
6.36±0.95
2.28±0.08
GJ 3470 b
13.67±1.59
3.88±0.33
GJ 436 b
22.25±2.22
4.17±0.17
776
612
578
599
692
642
HAT-P-11 b
25.75±2.86
HAT-P-26 b
18.75±2.22
K2-3 b
K2-19 c
HD 97658 b
HD 219134 b
4.46±0.47
7.63+0.95
−0.64
HIP 116454 b 11.76±1.27
15.90+7.70
−2.80
8.40±2.10
2.10+2.10
−1.30
11.10±3.50
16.30+6.00
−6.10
3.72±0.42
Kepler-10 b
K2-56 b
K2-3 d
K2-3 c
Kepler-10 c
13.98±1.79
Kepler-100 b
7.31±3.18
866
994
4.74±0.16
6.34+0.81
−0.36
1.61±0.09 1022
2.24+0.10
−0.09
2.54±0.18
758
691
500
371
783
4.51±0.47
2.08+0.18
−0.09
1.65+0.16
−0.07
1.53±0.11
2.23+0.14
−0.11
1.48+0.04
−0.03
2.36+0.09
−0.04
1.32±0.04 1271
2129
570
Kepler-102 d
3.81±1.91
1.18±0.04
Kepler-102 e
8.90±1.91
2.22±0.07
701
604
Kepler-103 b
9.85±8.58
Kepler-104 b 19.60+14.50
−12.40
3.70±2.00
Kepler-105 b
946
3.38±0.09
3.13+0.52
−0.70
2.22±0.11 1088
1043
993
863
583
931
859
714
636
562
844
714
628
613
Kepler-105 c
4.60±0.90
1.31±0.07
Kepler-106 c
10.49±3.18
2.50±0.33
Kepler-106 e
Kepler-11 b
Kepler-11 c
Kepler-11 d
Kepler-11 e
Kepler-11 f
11.12±5.72
1.91+1.27
−0.95
2.86+2.86
−1.59
7.31+0.95
−1.59
7.95+1.59
−2.22
1.91±0.95
Kepler-113 b 11.76±4.13
2.56±0.33
1.81+0.03
−0.04
2.87+0.04
−0.06
3.12+0.06
−0.07
4.20+0.07
−0.09
2.49+0.04
−0.07
1.82±0.04
Kepler-114 c
2.86±0.64
1.60±0.18
Kepler-120 b
Kepler-114 d
3.81±1.59
8.50+9.70
−7.50
27.70+11.40
−9.80
Kepler-131 b 16.21±3.50
Kepler-122 e
Kepler-131 c
8.26±6.04
Kepler-136 b 19.80+11.70
−10.40
0.07+0.06
Kepler-138 b
−0.04
1.97+1.91
−1.12
0.64+0.67
−0.39
Kepler-138 d
Kepler-138 c
2.54±0.28
2.31±0.37
2.02+0.70
−0.19
2.41±0.20
0.84±0.07
1.80+0.35
−0.17
0.53±0.03
1.20±0.07
1.21±0.08
ρp
a Ms Age Ωrot
FXUV
g cm-3 AU M⊙ Gyr Ω⊙ erg s-1 cm-2
Lhy
s−1
Len
s−1
Lhy/Len
Ref.a,b
5.14 0.015 0.91 10.2
0.57 0.092 1.10
3.3
1.28 0.098 0.91 11.0
7.97 0.017 0.91
0.63 0.154 1.12
0.31 0.247 1.12
5.79 0.015 0.18
2.97 0.014 0.18
1.29 0.031 0.51
1.69 0.029 0.47
1.34 0.053 0.81
0.41 0.048 0.82
1.3
5.0
5.0
5.0
6.0
2.5
6.0
6.5
9.0
5.94 0.038 0.79 12.5
3.72 0.080 0.77
3.98 0.091 0.78
0.96 0.100 0.95
5.18 0.077 0.61
2.58 0.141 0.61
17.22 0.209 0.61
8.11 0.241 0.96
6.31 0.017 0.91
5.89 0.240 0.91
17.37 0.073 1.08
12.86 0.086 0.81
4.47 0.117 0.81
1.41 0.128 1.09
3.53 0.094 0.81
1.87 0.060 0.96
11.28 0.072 0.96
3.69 0.111 1.00
3.66 0.243 1.00
1.78 0.091 0.96
0.67 0.107 0.96
1.33 0.155 0.96
0.59 0.195 0.96
0.68 0.250 0.96
10.80 0.050 0.75
3.82 0.065 0.56
1.29 0.083 0.56
3.80 0.055 0.59
6.0
2.0
· · ·
2.0
2.0
2.0
3.3
9.1
9.1
6.9
4.1
4.1
6.0
3.7
3.5
3.5
3.3
3.3
6.9
6.9
6.9
6.9
6.9
3.2
2.7
2.7
3.6
3.9
3.7
3.7
2.8
4.7
4.7
4.7
1.4
1.9
1.3
2.3
0.4
0.4
5.2
5.2
· · ·
0.3
1.1
1.2
0.3
0.4
1.5
0.5
· · ·
· · ·
· · ·
1.9
1.0
1.0
1.3
0.4
0.4
0.9
· · ·
1.6
1.6
0.2
0.2
0.2
0.2
0.2
0.2
0.2
0.3
0.9
0.9
· · ·
· · ·
0.2
0.2
· · ·
1.4
1.4
1.4
· · ·
173333.5
10959.0 3.4×1035 1.4×1035
2429.9
· · ·
· · ·
· · ·
293392.3
231.6
90.0
156736.6
181777.4
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
5353.9
699.6 1.3×1033 8.9×1033
5944.9 2.7×1034 2.4×1034
12237.8 4.4×1035 2.4×1035
926.6
485.1
· · ·
· · ·
· · ·
· · ·
· · ·
4879.7
540.3 6.4×1033 3.2×1033
1232.2
· · ·
· · ·
· · ·
374.9
170.1
1517.7
124224.8
639.6
14282.2
302.8
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
166.5
2165.8 3.1×1034 9.0×1033
5720.3
17615.2 3.8×1035 2.1×1035
12198.6
· · ·
· · ·
· · ·
· · ·
68.5
· · ·
· · ·
· · ·
· · ·
14.3
175.8 2.2×1035 8.7×1033
127.2 4.0×1035 1.5×1034
60.6 3.6×1032 3.5×1032
38.3 2.0×1033 4.9×1032
23.3 8.4×1034 2.8×1033
635.2
· · ·
· · ·
· · ·
4657.9
2796.6 6.1×1034 1.8×1034
1524.5
· · ·
· · ·
· · ·
1104.9
79.2
42.9
12210.7
2705.8
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
1838.5
924.8 8.4×1034 1.2×1034
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
De11R
· · ·
2.3 Mou14R
· · · Alo14R
Ba14R
· · ·
Pet15R
Pet15R
BT15R
Ha13R
Bi14R
· · ·
0.1 Mac14R
1.1 Bak10R
Ha11R
1.8
· · · Mot15R
· · · VG14R
V15R
· · ·
2.0 Ba15R,T
· · · Alm15R
· · · Alm15R
· · · Alm15R
E16R
· · ·
· · · W16R
· · · W16R
· · · Mar14R
· · · Mar14R
· · · Mar14R
3.4 Mar14R
HL14T
· · ·
1.8 JH15bT
JH15bT
· · ·
· · · Mar14R
· · · Mar14R
L13T
25.3
L13T
L13T
L13T
L13T
30.0
· · · Mar14R
X14T
· · ·
X14T
HL14T
HL14T
· · ·
· · · Mar14R
· · · Mar14R
HL14T
· · ·
JH15aT
JH15aT
· · ·
7.1 JH15aT
26.7
3.4
1.0
4.1
· · ·
· · ·
MNRAS 000, 1–9 (2016)
676 153.8
18.53 0.227 0.99
785
64.9
6.37 0.126 1.02
673 110.6
76.46 0.171 1.02
1060
78.6
18.72 0.106 1.20
449
408
343
2.1
30.5
11.6
2.51 0.075 0.52
6.28 0.090 0.52
1.98 0.128 0.52
Table A1 – continued Sub-Neptune Planet Parameters.
An overabundance of low-density Neptunes
11
ρp
FXUV
g cm−3 AU M⊙ Gyr Ω⊙ erg s-1 cm-2
a Ms Age Ωrot
Λ
50.1
56.6
91.4
79.2
7.6
13.7
20.7
23.9
22.6
49.4
9.28 0.054 0.77
9.95 0.068 0.77
7.56 0.102 0.83
5.56 0.163 0.83
0.43 0.222 1.07
0.12 0.270 1.07
4.7
4.7
4.7
4.7
4.4
4.4
4.84 0.045 0.97 10.7
0.57 0.075 0.97 10.7
0.27 0.117 0.97 10.7
6.55 0.141 0.87
990
793
693
1197
836
898
590 121.5
9.06 0.137 0.84
32.0
28.5
47.8
63.1
15.71 0.090 0.92
6.82 0.045 0.91
3.08 0.093 0.91
6.86 0.062 0.97
652 117.1
10.16 0.185 0.77
652
942
88.4
45.3
6.66 0.272 1.15
4.33 0.059 0.72
1108 105.3
34.86 0.047 0.86
47.5
61.1
44.1
26.5
79.8
77.6
12.22 0.075 1.11
9.12 0.124 1.11
4.48 0.074 0.51
2.51 0.114 0.51
9.24 0.272 1.06
6.71 0.140 0.88
993
462
372
635
640
489 101.1
3.28 0.202 0.80
20.6
34.8
13.7
29.9
41.6
49.0
64.4
75.4
30.6
45.9
40.8
23.2
11.7
12.0
36.7
56.4
23.8
17.2
8.5
5.2
72.7
53.2
65.1
48.5
25.3
2.65 0.070 1.19
0.96 0.113 1.19
1.93 0.105 0.97
1.31 0.085 0.54
1.70 0.107 0.54
0.32 0.191 0.65
3.71 0.203 1.10
4.05 0.269 1.10
1.93 0.062 0.75
2.83 0.081 0.75
4.03 0.210 1.08
1.18 0.330 1.08
0.66 0.092 0.98
0.71 0.109 0.98
1.05 0.186 0.99
0.19 0.534 0.99
1.24 0.056 0.76
0.93 0.073 0.76
0.54 0.093 0.98
0.40 0.108 0.98
1.56 0.260 1.21
4.55 0.050 0.54
5.15 0.090 0.54
4.42 0.051 0.98
2.72 0.066 0.98
Name
Mp
M⊕
Rp
R⊕
Teq
K
948
Kepler-161 b 12.10+7.40
−6.30
Kepler-161 c 11.80+10.50
−7.50
Kepler-176 c 23.00+13.50
−8.00
Kepler-176 d 15.20+10.40
−5.80
1.91+0.64
Kepler-177 b
−0.32
7.63+3.50
−3.18
6.99±3.50 2.00±0.10 1284
1.93+0.31
−0.14
1.87+0.30
−0.14
2.56+0.93
−0.26
2.47+0.90
−0.25
2.91+1.53
−0.30
7.10+3.71
−0.72
Kepler-177 c
Kepler-18 b
655
594
589
745
844
1.90±0.06 1277
1.23±0.04 1021
Kepler-18 c
17.16±1.91 5.50±0.26
−0.34
−1.09
−0.18
Kepler-20 b
Kepler-18 d
16.53±1.27 6.99±0.33
1.95+0.84
8.80+7.40
Kepler-184 c
−0.19
−5.70
−10.60 2.40+0.52
Kepler-189 c 22.70+17.10
5.30+3.30
Kepler-197 c
−2.90
8.58±2.22 1.91+0.12
−0.21
3.07+0.20
16.21+3.18
Kepler-20 c
−0.31
−3.81
Kepler-211 c 18.30+22.40
−17.00 2.45+1.62
−11.90 2.34+0.44
Kepler-215 d 23.60+17.30
−17.70 2.51+0.63
Kepler-219 d 19.10+29.90
−0.40
2.68+0.61
Kepler-221 c 15.10+10.10
−0.26
−6.70
−10.10 1.56+0.58
Kepler-226 b 24.00+11.80
15.20+3.20
Kepler-23 b
−2.90
17.60+13.70
−11.90 2.20±0.07
1.82+0.26
4.90+1.80
−0.25
−1.30
2.20+1.50
1.69+0.24
Kepler-231 c
−0.23
−1.10
2.00+0.85
Kepler-238 f 13.35+2.86
−0.17
−2.54
−13.00 2.32+0.83
Kepler-244 d 15.20+20.00
−12.10 3.31+1.40
Kepler-245 d 21.60+16.60
Kepler-25 b
Kepler-231 b
Kepler-23 d
−0.18
−0.37
−0.12
9.60±4.20 2.71±0.05 1305
Kepler-25 c
Kepler-254 c
Kepler-26 b
24.60±5.70 5.21±0.09 1029
2.09+0.84
3.20+3.70
−0.16
−2.70
5.10±0.70 2.78±0.11
845
465
Kepler-26 c
6.20±0.70 2.72±0.12
7.17+0.38
21.20+3.20
Kepler-27 c
−0.27
−3.70
2.91+1.27
Kepler-276 c 16.53+4.45
−0.28
−3.50
2.81+1.23
Kepler-276 d 16.21+5.09
−0.27
−4.45
8.80+3.80
Kepler-28 b
2.93±0.46
−3.10
10.90+6.10
−4.50
7.31±6.67 2.15±0.10
Kepler-289 b
Kepler-28 c
2.77±0.44
Kepler-289 d
4.13±0.95 2.68±0.17
Kepler-29 b
4.50±1.50 3.35±0.22
Kepler-29 c
4.00±1.30 3.14±0.20
Kepler-30 b
11.30±1.40 3.90±0.20
Kepler-307 c
Kepler-307 b
23.10±2.70 8.79±0.50
Kepler-30 d
3.60+0.90
Kepler-305 b 10.49+2.54
−0.36
−1.91
6.04+2.54
3.30+0.82
Kepler-305 c
−2.22
−0.33
3.18±0.64 3.20+1.20
−0.46
2.81+1.05
1.59+0.64
−0.42
−0.32
29.50+9.60
4.71+2.23
−7.00
−0.57
9.40+3.60
2.25±0.11
−3.10
7.70+5.00
Kepler-32 c
−3.80
Kepler-326 c 17.40+15.40
6.90+8.50
Kepler-326 d
−5.90
2.02±0.11
−10.70 2.79+1.96
−1.31
2.41+1.70
−1.13
Kepler-32 b
Kepler-31 c
MNRAS 000, 1–9 (2016)
415
457
669
581
743
650
630
502
872
802
599
353
927
808
883
818
653
595
443
974
856
Lhy
s−1
· · ·
· · ·
· · ·
Len
s−1
· · ·
· · ·
· · ·
7681.3
4844.0
1567.5
· · ·
613.8
962.8 4.1×1036 2.0×1035
651.5 5.8×1035 1.0×1035
· · ·
· · ·
· · ·
46258.8
16344.6 3.3×1035 2.1×1035
6729.1 3.5×1035 1.9×1035
1017.7
· · ·
· · ·
497.7
3777.0
732.8
174.6
12588.9
2352.6
1158.9
70313.6
7076.1
8385.5
3067.7
2185.7
921.0
507.0
720.1
987.3
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
92614.7
35845.4 1.6×1035 1.7×1035
2984.8 9.1×1034 2.9×1034
5260.6 2.7×1034 2.8×1034
3338.5 5.8×1033 1.7×1034
222.8
· · ·
· · ·
1237.2
704.2
7830.8
4587.9
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
4490.4
1818.4 4.1×1034 1.9×1034
13666.1 1.3×1036 4.0×1035
9778.1 8.9×1035 2.8×1035
718.8 7.5×1033 4.3×1033
86.7
· · ·
· · ·
3742.0 5.1×1034 2.4×1034
2158.9 8.0×1034 2.5×1034
11054.6 3.9×1036 9.5×1035
8159.4 9.8×1037 1.2×1036
1108.1
· · ·
· · ·
2232.3
689.0
· · ·
· · ·
· · ·
· · ·
34240.8
20445.5 4.7×1034 6.1×1034
· · ·
· · ·
Lhy/Len
Ref.a,b
20.5
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
HL14T
HL14T
HL14T
HL14T
X14T
X14T
5.8
· · · Co11R,T
1.6 Co11R,T
1.8 Co11R,T
HL14T
· · ·
HL14T
HL14T
Ga12R
Ga12R
HL14T
HL14T
HL14T
HL14T
HL14T
HL14T
HL14T
K14T
K14T
X14T
HL14T
HL14T
· · ·
· · · Mar14R
0.9 Mar14R
3.1 HL14T
1.0 JH15bT
0.3 JH15bT
HL14T
· · ·
X14T
X14T
HL14T
HL14T
S14T
S14T
2.2
3.3 JH15bT
3.2 JH15bT
SO12T
1.7
SO12T
X14T
X14T
X14T
X14T
HL14T
HL14T
HL14T
HL14T
· · ·
0.8 HL14T
3.2
2.1
4.1
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
81.7
1.0
1.0
0.8
0.8
· · ·
· · ·
1.9
1.9
1.9
· · ·
· · ·
· · ·
0.2
0.2
2.0
· · ·
1.6
2.9
· · ·
· · ·
· · ·
1.2
1.2
· · ·
· · ·
1.4
3.9
3.9
1.1
2.0
2.0
0.5
· · ·
· · ·
1.5
1.5
3.0
3.0
2.3
2.3
1.1
1.1
· · ·
· · ·
2.0
2.0
· · ·
0.8
0.8
2.8
2.8
4.5
5.2
5.4
5.2
5.2
1.6
1.6
4.6
4.6
4.4
6.6
6.6
3.5
3.5
6.8
4.8
3.6
2.9
2.9
4.0
3.0
3.0
1.6
3.8
3.8
2.2
2.2
3.8
3.8
· · ·
· · ·
1.6
1.6
4.4
4.4
6.6
6.6
4.2
2.7
2.7
4.7
4.7
12
P. Cubillos et al.
Table A1 – continued Sub-Neptune Planet Parameters.
Name
Mp
M⊕
Rp
R⊕
Teq
K
Λ
ρp
FXUV
g cm−3 AU M⊙ Gyr Ω⊙ erg s-1 cm-2
a Ms Age Ωrot
Lhy
s−1
Len
s−1
Lhy/Len Ref.a,b
−0.24
4.03±0.38
Kepler-33 e
Kepler-33 d
941
830
762
3.20±0.30 1113
−17.70 1.18±0.11
−12.40 2.30+0.97
Kepler-327 c 20.30+27.10
Kepler-328 b 28.61+13.03
0.80+2.50
Kepler-33 c
−0.70
4.70±2.00 5.35±0.49
6.70+1.20
−1.30
11.50+1.80
Kepler-33 f
4.47±0.42
−2.10
Kepler-333 b 28.20+29.00
−24.10 1.61±0.17
8.50+7.20
Kepler-338 e
−6.30
7.30+7.80
Kepler-339 c
−6.20
Kepler-339 d 14.70+14.10
Kepler-350 c
Kepler-350 d 14.94+5.40
−4.77
4.80+5.70
Kepler-351 b
−4.70
Kepler-351 c 11.10+9.90
−7.60
4.46+0.33
Kepler-36 b
−0.27
8.10+0.60
Kepler-36 c
−0.46
Kepler-385 b 12.80+11.40
−6.80
Kepler-385 c 13.20+16.80
−9.00
Kepler-396 c 17.80+2.86
−1.27
Kepler-4 b
1.56±0.07 1258
1.16+0.45
−0.09
−10.00 1.18+0.46
−0.09
6.04±3.18 3.11+1.43
−0.61
2.81+1.28
−0.54
3.03+1.33
−0.24
3.16+1.38
−0.25
1.48±0.03 1069
3.68±0.05 1014
2.79+1.62
−0.35
3.10+1.81
−0.39
5.31+1.95
−0.99
24.47±3.81 4.01±0.21 1614
1009
1041
924
888
542
468
909
440
Kepler-406 b
6.36±1.27 1.44±0.03 1452
Kepler-406 c
2.86±1.91 0.85±0.03 1171
Kepler-454 b
6.84±1.40 2.37±0.13
916
Kepler-48 b
3.94±2.10 1.92±0.10 1024
Kepler-48 c
14.61±2.30 2.71±0.14
809
Kepler-60 c
3.90±0.80 1.90±0.15 1092
Kepler-48 d
Kepler-49 b
Kepler-49 c
Kepler-51 b
Kepler-51 c
Kepler-51 d
Kepler-54 b
Kepler-54 c
Kepler-56 b
Kepler-57 c
Kepler-60 b
Kepler-60 d
Kepler-65 c
Kepler-68 b
Kepler-68 c
Kepler-78 b
Kepler-79 b
Kepler-79 c
Kepler-79 d
Kepler-79 e
Kepler-80 b
Kepler-80 c
Kepler-81 b
Kepler-81 c
Kepler-82 c
561
492
546
627
2.72±0.12
2.55±0.13
7.95±4.61 2.04±0.11
7.80+15.40
−3.90
7.90+15.60
−3.90
2.22+1.59
7.10±0.30
−0.95
4.13±0.32 9.01+2.81
−1.71
7.63±0.95 9.71±0.50
21.50+5.60
−4.90
19.80+6.60
−4.40
22.25+3.81
−3.50
9.30+25.20
−3.00
4.20±0.60 1.71±0.13 1177
1.32±0.08
6.52+0.29
−0.28
1.55±0.67
2.19±0.07
1496
453
394
705
992
−18.50 2.61±0.04 1363
1252
4.20±0.80 1.99±0.16
26.60+20.40
2.31+0.06
8.26+2.22
−0.09
−2.54
4.77+2.54
0.95+0.03
−0.04
−3.50
1.91±0.32 1.18+0.16
−0.09
10.90+7.31
3.48±0.07
−6.04
6.04+1.91
−2.29
6.04+2.10
−1.59
4.13+1.21
−1.11
5.70+7.40
−4.10
7.00+8.90
−5.50
16.40+27.00
−9.90
4.00+6.70
−2.40
19.10+55.50
−4.90
3.73±0.08
7.17+0.13
−0.16
3.49±0.13
2.65±0.11
2.79±0.13
2.42±0.38
2.37±0.37
5.35±2.45
1033
2217
992
784
633
546
700
633
707
559
500
591 220.6
68.14 0.047 0.55
627 150.3
12.96 0.219 1.15
1.7
7.1
15.2
25.6
0.13 0.119 1.29
0.17 0.166 1.29
0.57 0.214 1.29
0.71 0.254 1.29
506 262.1
37.27 0.087 0.54
32.8
51.6
12.35 0.090 1.10
25.79 0.069 0.84
804 117.4
49.34 0.091 0.84
1.11 0.134 1.00
3.73 0.172 1.00
0.95 0.214 0.89
1.94 0.287 0.89
7.52 0.115 1.07
0.90 0.128 1.07
3.25 0.097 1.09
2.44 0.127 1.09
0.66 0.368 0.85
2.10 0.046 1.22
11.83 0.036 1.07
25.43 0.056 1.07
2.83 0.095 1.03
3.07 0.053 0.88
4.05 0.085 0.88
5.14 0.230 0.88
2.14 0.060 0.55
2.63 0.079 0.55
0.03 0.251 1.04
0.03 0.384 1.04
11.29 0.063 0.51
47.48 0.082 0.51
0.44 0.103 1.32
13.77 0.092 0.76
4.63 0.073 1.04
3.14 0.085 1.04
2.94 0.103 1.04
8.25 0.068 1.25
3.69 0.062 1.08
30.30 0.091 1.08
6.43 0.009 0.76
1.43 0.117 1.17
0.64 0.187 1.17
0.09 0.287 1.17
0.54 0.386 1.17
1.69 0.065 0.72
1.77 0.079 0.72
6.35 0.055 0.64
1.66 0.089 0.64
0.69 0.257 0.85
14.6
45.4
22.2
56.9
21.3
16.5
33.4
35.5
57.7
28.7
23.1
21.7
23.9
15.2
50.5
59.9
34.7
43.0
4.2
7.7
17.3
64.5
15.8
14.2
16.1
56.7
21.7
36.7
5.5
24.0
15.7
10.1
16.4
23.3
30.0
72.5
22.9
54.0
15.1
0.05 0.509 1.04
605 122.9
530 214.3
3.0
2.6
4.4
4.4
4.4
4.4
3.2
4.8
4.4
4.4
3.2
3.2
4.9
4.9
4.8
4.8
3.3
3.3
4.4
6.8
2.1
2.1
4.3
6.2
6.2
6.2
2.9
2.9
3.4
3.4
3.4
4.2
4.2
4.4
4.7
5.1
5.1
5.1
3.3
6.3
6.3
0.8
3.2
3.2
3.2
3.2
2.9
2.9
3.6
3.6
4.7
1.1
· · ·
1.0
1.0
1.0
1.0
1.1
· · ·
1.6
1.6
8.5
8.5
· · ·
· · ·
1.6
1.6
· · ·
· · ·
2.0
0.8
0.2
0.2
1.2
0.3
0.3
0.3
1.5
1.5
3.3
3.3
3.3
0.8
0.8
· · ·
0.8
· · ·
· · ·
· · ·
4.0
0.2
0.2
1.7
· · ·
· · ·
· · ·
· · ·
1.1
1.1
2.0
2.0
· · ·
243252.4 1.6×1036 1.8×1036
146470.2
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
321.0
758.8
759.5
296.8
9535.9
4731.3
4051.6
1032.9
7871.5
4264.3
10799.4
13691.2
356.0 5.6×1033 3.6×1033
197.9
10334.7
8348.0 2.5×1035 1.2×1035
8110.4
· · · HL14T
X14T
1487.0
· · ·
2707.6 2.2×1042 1.4×1036 1.6×106 HL16T
24.0 HL16T
1385.8 1.8×1037 7.5×1035
1.2 HL16T
837.4 3.3×1034 2.8×1034
3.0 HL16T
595.6 1.8×1034 6.1×1033
· · · HL14T
2381.1
· · · HL14T
· · · HL14T
· · · HL14T
X14T
0.9
X14T
· · ·
1.6 HL14T
· · · HL14T
· · · Ca12T
2.1 Ca12T
· · · HL14T
· · · HL14T
X14T
· · ·
Bo10R
· · ·
· · · Mar14R
· · · Mar14R
· · · Ge15R
· · · Mar14R
· · · Mar14R
· · · Mar14R
X13T
· · ·
X13T
180.0 Mas14T
11.8 Mas14T
3.4 Mas14T
· · · HL14T
· · · HL14T
H13T
1.9
X13T
· · ·
· · · JH15bT
· · · JH15bT
· · · JH15bT
· · · HL14T
Gi13R
· · ·
Gi13R
· · ·
· · · Pep13R
JH14T
1.2
JH14T
JH14T
JH14T
X13T
X13T
X13T
X13T
X13T
6920.5 7.6×1034 6.3×1034
2709.1 1.7×1035 7.4×1034
1150.1 1.3×1037 6.1×1035
635.8 6.8×1034 2.0×1034
3568.7 3.6×1034 1.6×1034
2391.0 1.3×1034 9.6×1033
17772.7
6948.7 6.6×1034 3.0×1034
265.2
5074.4
3724.8 1.8×1039 1.0×1037
1596.5 5.3×1037 4.5×1036
908.7 1.3×1036 3.8×1035
2727.7
1610.1
18720.8 7.8×1035 4.2×1035
1761.3
96951.2
5999.3
4082.8
8833.6
8080.2
441.3
204.7
21.3
40.7
2.3
3.4
2.3
1.4
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.2
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
731021.4
· · ·
· · ·
MNRAS 000, 1–9 (2016)
Table A1 – continued Sub-Neptune Planet Parameters.
An overabundance of low-density Neptunes
13
Name
Mp
M⊕
Rp
R⊕
Teq
K
ρp
FXUV
g cm−3 AU M⊙ Gyr Ω⊙ erg s-1 cm-2
a Ms Age Ωrot
2.37±0.35
7.50+13.80
Kepler-83 c
485
−3.60
Kepler-84 b 21.20+32.00
−13.80 2.23±0.10 1092
Kepler-85 b 15.30+22.10
−10.80 1.97±0.10
884
Kepler-85 c 24.00+34.40
−16.70 2.18±0.10
6.40±0.80 6.15±0.09
Kepler-87 c
8.58±2.54 3.78+0.39
−0.36
Kepler-88 b
444
801
Kepler-89 b 10.49±4.45 1.72±0.16 1624
771 108.3
9.40+2.40
Kepler-89 c
−2.10
Kepler-89 e 13.00+2.50
−2.10
Kepler-92 c
6.04±1.91 2.60±0.08
4.41±0.42 1154
6.70±0.63
665
856
Kepler-93 b
4.02±0.68 1.48±0.02 1138
Kepler-94 b 10.81±1.27 3.51±0.15 1095
Kepler-95 b 13.03±2.86 3.42±0.09 1019
Kepler-96 b
8.58±3.50 2.67±0.22
782
Kepler-97 b
3.50±1.91 1.48±0.13 1451
Kepler-98 b
3.50±1.59 2.00±0.22 1743
Kepler-99 b
6.15±1.30 1.48±0.08
WASP-47 d 15.20±7.00 3.64±0.13
880
983
Λ
49.5
65.9
66.4
17.7
21.5
28.5
14.0
22.1
20.5
18.1
21.3
28.3
31.2
12.3
7.6
35.8
32.2
3.12 0.126 0.66
10.50 0.083 1.00
10.95 0.078 0.92
12.83 0.103 0.92
0.15 0.671 1.10
0.88 0.095 0.96
11.43 0.051 1.28
0.60 0.101 1.28
0.24 0.305 1.28
1.89 0.186 1.21
6.82 0.053 0.91
1.38 0.034 0.81
1.79 0.102 1.08
2.48 0.125 1.00
5.93 0.036 0.94
2.42 0.026 0.99
10.46 0.050 0.79
3.1
9.6
4.0
4.0
7.2
2.2
3.3
3.3
3.3
5.9
5.2
2.5
8.7
3.7
4.6
8.3
4.3
1.74 0.086 1.04
· · ·
1.8
1.4
1.2
1.2
1.3
1.1
2.6
2.6
2.6
1.7
0.3
0.4
0.3
0.3
0.3
0.2
0.4
1.4
Lhy
s−1
· · ·
· · ·
· · ·
Len
s−1
· · ·
· · ·
· · ·
2880.2
11252.1
4726.9
· · ·
· · ·
· · ·
· · ·
2738.9
178.9 9.6×1034 2.5×1034
2566.9 6.2×1034 2.2×1034
78217.2
20126.4 7.7×1035 4.3×1035
2227.5 1.7×1035 7.7×1034
3067.7 4.3×1034 1.6×1034
942.4
1815.1 3.5×1034 1.2×1034
281.0 2.6×1033 2.8×1033
128.8
· · ·
· · ·
· · ·
· · ·
1966.4
2774.0
· · ·
· · ·
· · ·
· · ·
883.3
6812.1 3.6×1034 2.8×1034
· · ·
· · ·
Lhy/Len Ref.a,b
· · ·
· · ·
· · ·
· · ·
3.8
X13T
X13T
X13T
X13T
O14T
N13T
2.8
· · · W13R
1.8 Mas13T
2.2 Mas13T
X14T
2.8
· · · Dr15R
2.9 Mar14R
0.9 Mar14R
· · · Mar14R
· · · Mar14R
· · · Mar14R
· · · Mar14R
1.3 Be15T
a References for Planetary Masses and Radii. Alm15: Almenara et al. (2015), Alo14: Alonso et al. (2014), BT15: Berta-Thompson et al.
(2015), Ba14: Barros et al. (2014), Ba15: Barros et al. (2015), Bak10: Bakos et al. (2010), Be15: Becker et al. (2015), Bi14: Biddle et al.
(2014), Bo10: Borucki et al. (2010), Ca12: Carter et al. (2012), Co11: Cochran et al. (2011), De11: Demory et al. (2011), Dr15:
Dressing et al. (2015), E16: Espinoza et al. (2016), Ga12: Gautier et al. (2012), Ge15: Gettel et al. (2016), Gi13: Gilliland et al. (2013),
H13: Huber et al. (2013), HL14: Hadden & Lithwick (2014), HL16: Hadden & Lithwick (2016), Ha11: Hartman et al. (2011), Ha13:
Harpsøe et al. (2013), JH14: Jontof-Hutter et al. (2014), JH15a: Jontof-Hutter et al. (2015), JH15b: Jontof-Hutter et al. (2016), K14:
Kipping et al. (2014), L13: Lissauer et al. (2013), Mac14: Maciejewski et al. (2014), Mar14: Marcy et al. (2014), Mas13: Masuda et al.
(2013), Mas14: Masuda (2014), Mot15: Motalebi et al. (2015), Mou14: Moutou et al. (2014), N13: Nesvorn´y et al. (2013), O14: Ofir et al.
(2014), Pep13: Pepe et al. (2013), Pet15: Petigura et al. (2016), S14: Schmitt et al. (2014), SO12: Sanchis-Ojeda et al. (2012), V15:
Vanderburg et al. (2015), VG14: Van Grootel et al. (2014), W13: Weiss et al. (2013), W16: Weiss et al. (2016), X13: Xie (2013), X14:
Xie (2014).
b The 'R' and 'T' superscripts indicate RV- and TTV-estimated mass, respectively.
MNRAS 000, 1–9 (2016)
|
1604.05191 | 2 | 1604 | 2016-05-13T12:03:37 | Giant planet formation in radially structured protoplanetary discs | [
"astro-ph.EP"
] | Our recent N-body simulations of planetary system formation, incorporating models for the main physical processes thought to be important during the building of planets (i.e. gas disc evolution, migration, planetesimal/boulder accretion, gas accretion onto cores, etc.), have been successful in reproducing some of the broad features of the observed exoplanet population (e.g. compact systems of low mass planets, hot Jupiters), but fail completely to form any surviving cold Jupiters. The primary reason for this failure is rapid inward migration of growing protoplanets during the gas accretion phase, resulting in the delivery of these bodies onto orbits close to the star. Here, we present the results of simulations that examine the formation of gas giant planets in protoplanetary discs that are radially structured due to spatial and temporal variations in the effective viscous stresses, and show that such a model results in the formation of a population of cold gas giants. Furthermore, when combined with models for disc photoevaporation and a central magnetospheric cavity, the simulations reproduce the well-known hot-Jupiter/cold-Jupiter dichotomy in the observed period distribution of giant exoplanets, with a period valley between 10-100 days. | astro-ph.EP | astro-ph | MNRAS 000, 000–000 (0000)
Preprint 16 May 2016
Compiled using MNRAS LATEX style file v3.0
Giant planet formation in radially structured
protoplanetary discs
Gavin A. L. Coleman(cid:63) & Richard P. Nelson
Astronomy Unit, Queen Mary University of London, Mile End Road, London, E1 4NS, U.K.
6
1
0
2
y
a
M
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
9
1
5
0
.
4
0
6
1
:
v
i
X
r
a
ABSTRACT
Our recent N-body simulations of planetary system formation, incorporating models
for the main physical processes thought to be important during the building of plan-
ets (i.e. gas disc evolution, migration, planetesimal/boulder accretion, gas accretion
onto cores, etc.), have been successful in reproducing some of the broad features of
the observed exoplanet population (e.g. compact systems of low mass planets, hot
Jupiters), but fail completely to form any surviving cold Jupiters. The primary rea-
son for this failure is rapid inward migration of growing protoplanets during the gas
accretion phase, resulting in the delivery of these bodies onto orbits close to the star.
Here, we present the results of simulations that examine the formation of gas giant
planets in protoplanetary discs that are radially structured due to spatial and tempo-
ral variations in the effective viscous stresses, and show that such a model results in
the formation of a population of cold gas giants. Furthermore, when combined with
models for disc photoevaporation and a central magnetospheric cavity, the simulations
reproduce the well-known hot-Jupiter/cold-Jupiter dichotomy in the observed period
distribution of giant exoplanets, with a period valley between 10–100 days.
Key words: planetary systems, planets and satellites: formation, planets-disc inter-
actions, protoplanetary discs.
1 INTRODUCTION
Ever since the discovery of the first extrasolar giant planet
around a main sequence star (Mayor & Queloz 1995), ques-
tions have been asked as to the formation and evolution
of giant exoplanets. To date over 1640 confirmed extraso-
lar planets have been discovered, displaying a broad range
of orbital and physical properties, and approximately 600 of
these are believed to be gas giants (Han et al. 2014). Explain-
ing the origins of the broad diversity of exoplanets remains
a formidable challenge to planet formation theory, and even
the more restricted challenge of explaining the orbital period
distribution of giant exoplanets has not yet been addressed
satisfactorily.
Observational biases, in particular the fact that ground
based transit surveys are only sensitive to detecting giant
planets with orbital periods (cid:46) 10 days, and that radial ve-
locity searches have surveyed stars that are more metal-rich
than the average, give the impression that hot Jupiters are
common. Recent studies that have examined data from the
Kepler spacecraft and follow-up radial velocity measure-
ments (Fressin et al. 2013; Santerne et al. 2015) find that
hot Jupiters are expected to orbit only 1% of main sequence
stars, while cold Jupiters have a higher occurrence rate of
(cid:63) Email: [email protected]
c(cid:13) 0000 The Authors
17% (Cassan et al. 2012). The occurrence rate between the
two populations does not increase linearly, however, as a 'pe-
riod valley' exists between 10–85 days where there is a dearth
of giant planet detections when accounting for observational
biases. This period valley was first observed in radial veloc-
ity surveys (Udry et al. 2003; Cumming et al. 2008), and
its existence has been supported by the aforementioned re-
cent analysis of combined Kepler and radial velocity obser-
vations (Santerne et al. 2015). Individual theories have been
put forward to explain this period valley (Hasegawa & Pu-
dritz 2011; Alexander & Pascucci 2012; Ercolano & Rosotti
2015), but none have been incorporated into ab initio models
of planet formation to examine whether or not it is possi-
ble to explain this, and other features in the giant planet
distribution, from first principles.
Competing theories of giant planet formation, including
the core-accretion and pebble-accretion models (e.g. Ida &
Lin 2004; Alibert et al. 2006; Mordasini et al. 2009; Bitsch
et al. 2015b), and the tidal-downsizing model (e.g. Nayak-
shin 2015), have been used to make predictions about the
giant planet population for comparison with observations,
and to examine the formation of giant planets in our So-
lar System (Levison et al. 2015). The fact that many mul-
tiplanet systems have been discovered, containing various
combinations of super-Earths, Neptunes and Jovian mass
bodies (Muirhead et al. 2012; Neveu-VanMalle et al. 2016;
2 G. A. L. Coleman and R. P. Nelson
Becker et al. 2015), often in compact systems (e.g. Lissauer
et al. 2011; Becker et al. 2015), suggests that gravitational
interactions, and perhaps competitive accretion, are essen-
tial components of the planet formation process. Further-
more, the fact that many giant planets appear to be on ec-
centric orbits suggests that dynamical instabilities involving
initially compact giant planet systems, either during or after
formation, are common and important for shaping planetary
system architectures (e.g. Rasio & Ford 1996). For this rea-
son, we have adopted the approach of using N-body simu-
lations, combined with prescriptions for the major physical
processes thought to occur during planet formation, to sim-
ulate the building and early evolution of planetary systems.
In recent work (Coleman & Nelson 2014, 2016), we have
examined the formation of planets in irradiated, viscous
disc models that have adopted the standard α prescription
(Shakura & Sunyaev 1973; Lynden-Bell & Pringle 1974). Ex-
cept for a small region close to the star where the temper-
ature exceeds 1000 K, we have assumed that α is constant,
leading to smooth temperature and surface density profiles
in the discs. The models have been successful in forming
systems containing hot Jupiters, multiple super-Earths and
Neptunes in compact configurations, and numerous terres-
trial planets with a variety of compositions, but the mod-
els fail completely to form any surviving cold Jupiters. The
main reason for this is that giant planet cores undergo rapid
inward migration as they accrete gas, because the corota-
tion torques that are needed to counteract the Lindblad
torques become saturated (e.g. Paardekooper et al. 2011).
These planets then end up as hot planets orbiting close to
the star. Coleman & Nelson (2014) undertook a detailed ex-
amination of the conditions required for giant planet forma-
tion and survival, and showed that a Jovian mass planet that
settles into a final orbit at 5 au must have initiated runaway
gas accretion and type II migration when at an orbital radius
∼ 15 au, and this should have occurred late in the disc life-
time so that the gas disc disperses before the planet type II
migrates all the way to the central star (or into the magneto-
spheric cavity if one is present). In this paper, we address the
question of whether or not radial structuring of a protoplan-
etary disc, because of spatial and temporal variations in the
viscous stress, can prevent accreting giant planet cores from
migrating inwards rapidly because of the 'planet traps' cre-
ated by the surface density variations (Masset et al. 2006).
Although we adopt a simple, proof-of-concept 'toy model'
for the generation of radial structuring of the disc, our re-
sults suggest that this may provide an effective means of
allowing the formation of surviving cold Jupiters, and point
the way to an avenue of potentially fruitful future research.
The paper is organised as follows. We describe the up-
dates to our physical model and numerical methods in sect.
2. We present our results in sect. 3, and draw our conclusions
in sect. 4.
2 PHYSICAL MODEL AND NUMERICAL
METHODS
The N-body simulations presented here were performed us-
ing the Mercury-6 symplectic integrator (Chambers 1999),
adapted to include the additional physical processes de-
scribed below. Some of these are updated versions of those
described in Coleman & Nelson (2016), and some of the pro-
cesses are new to this paper.
2.1 Recap of the physical model
The disc and planet formation models from Coleman & Nel-
son (2016) contain the following elements:
(i) The N-body component computes the gravitational inter-
action between numerous planetary embryos and a swarm of
orbiting planetesimals/boulders. The planetary embryos ex-
perience gravitational forces from each other and from the
planetesimals/boulders, whereas the planetesimals/boulders
experience gravitational forces from the embryos only, and
not from the other small bodies. The planetesimals/boulders
are treated as "super-planetesimals" that are supposed to
represent a large collection of actual small bodies. The
masses of these super-planetesimals are therefore much
larger than the mass of an individual small body, but also
significantly smaller than the embryo mass. The super-
planetesimals experience gas drag, and are therefore as-
signed a physical radius in the range 10 m ≤ R ≤ 10 km.
(ii) The gas disc is evolved by solving the diffusion equation
for a 1D viscous α-disc model (Shakura & Sunyaev 1973;
Lynden-Bell & Pringle 1974). Temperatures are calculated
by balancing viscous heating and stellar irradiation with
black-body cooling. We use two values of the background
α value in our simulations, 2 × 10−3 and 6 × 10−3. We as-
sume that solar metallicity corresponds to a gas-to-solids
ratio of 100:1, and that half of the grains in the disc remain
small and contribute to the opacity, and half of them grow to
form larger planet-building bodies. When a giant planet is
present, tidal torques from the planet are applied to the disc
leading to the opening of a gap (Lin & Papaloizou 1986).
(iii) We mimic the presence of fully-developed MHD turbu-
lence in the inner disc, when the disc temperature exceeds
1000 K (Umebayashi & Nakano 1988; Desch & Turner 2015),
by increasing the viscosity parameter there by a factor of 5
above the background values of α = 2 × 10−3 or 6 × 10−3.
(iv) We incorporate models for the photoevaporation of the
disc. We have implemented a standard photoevaporative
wind model (Dullemond et al. 2007), where the wind is as-
sumed to be launched thermally from the disc upper and
lower surfaces beyond a critical radius that corresponds ap-
proximately to the thermal velocity being equal to the es-
cape velocity. We have also implemented a direct photoe-
vaporation model that can be switched on, in addition to
the standard model, if the inner gas disc becomes depleted
and optically thin such that the stellar radiation acts di-
rectly on the disc inner edge that faces the star (Alexan-
der & Armitage 2009). The influence of including, or not
including, the direct photoevaporation component on the fi-
nal outcomes of the simulations is one of the issues that we
investigate in this paper, along with the effect of neglecting
photoevaporation altogether.
(v) Boulders and planetesimals orbiting in the disc expe-
rience aerodynamic drag, with the drag law (Stokes or Ep-
stein) depending on the objects size and the molecular mean
free path (Weidenschilling 1977). For reference, we note that
the drag-induced migration time scales associated with the
different sized bodies are as follows (assuming a disc mass
equivalent to the minimum mass solar nebula). A 10 m boul-
der migrates from 1 au into the central magnetospheric cav-
MNRAS 000, 000–000 (0000)
ity at 0.04 au in approximately 103 yr, and from 20 au in
approximately 106 yr. A 100 m planetesimal migrates from
1 au within approximately 5 × 105 yr, and within the disc
lifetime it migrates from 10 au to 6 au. The 1 km and 10 km
planetesimals display very small amounts of migration over
the disc lifetime.
(vi) Torque formulae from Paardekooper et al. (2010, 2011)
are used to simulate type I migration due to Lindblad and
corotation torques, and our model accounts for the possible
saturation of the corotation torque. The influences of eccen-
tricity and inclination on the disc forces are also included
(Cresswell & Nelson 2008; Fendyke & Nelson 2014).
(vii) Type II migration of gap forming planets is simu-
lated using the impulse approximation of Lin & Papaloizou
(1986), where we use the gap opening criterion of Crida et al.
(2006) to determine when to switch between type I and II
migration. Thus, when a planet is in the gap opening regime,
the planet exerts tidal torques on the disc to open a gap, and
the disc back-reacts onto the planet to drive type II migra-
tion in a self-consistent manner.
(viii) The accretion of gaseous envelopes on to solid cores
occurs once a planet's mass exceeds 3 M⊕, and the gas ac-
cretion rate is obtained from analytic fits to the detailed 1D
models based on solving the planetary structure equations
in the presence of prescribed rates of planetesimal accretion
presented by Movshovitz et al. (2010). Gas that is accreted
onto a planet is removed from the surrounding disc, such
that the accretion scheme conserves mass.
(ix) The effective capture radius of a protoplanet that ac-
cretes boulders/planetesimals is enhanced by atmospheric
drag using the prescription from Inaba & Ikoma (2003).
We adopt an inner boundary to our simulation domain
at 0.04 au, which we assume to represent the outer edge of
an inner magnetospheric cavity. Any planet that enters this
region no longer evolves, unless another planet enters the
cavity, in which case the latter body is retained and the for-
mer one is assumed to have been pushed into the star. This
is repeated for all subsequent planets that pass through the
inner boundary (note that no sub-Neptune mass planets en-
tered the cavity and pushed any giants into the star). When
presenting our results in Figs. 6, 8, 9 and 10, we reassign the
final semimajor axes of these inner planets to straddle the
stopping radius at 0.04 au, in order to mimic our expecta-
tion that the inner cavities will have a range of radii. This
reassignment assumes that the distribution of cavity edges
is Gaussian with standard deviation of 0.01 au.
2.2 Model improvements
2.2.1 Gas envelope accretion
A planet undergoes runaway gas accretion once the enve-
lope and core are of comparable mass, and during this phase
the planet rapidly accretes the material occupying its feed-
ing zone, until it reaches its 'gas isolation mass', where the
feeding zone is now empty and a gap has formed in the
disc. Coleman & Nelson (2014) obtained fits to the runaway
gas accretion rates from 2D hydrodynamic simulations, but
they only considered the migration of the rapidly accret-
ing planet after it had reached its 'gas isolation mass'. We
have improved on the fits of Coleman & Nelson (2014) by
allowing the planet to migrate while undergoing runaway
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
3
gas accretion. Including migration in the determination of
the fits makes them more consistent with the hydrodynamic
models.
Once a planet enters the runaway gas accretion phase
prior to reaching the gap forming mass, we apply the follow-
ing steps:
(i) Calculate the gas isolation mass, miso, according to:
miso = 2πrpΣg(rp)∆r
(1)
where Σg(rp) is the gas surface density taken at the planet's
location, and ∆r is given by
√
∆r = 6
3RH
(2)
where RH is the planet's Hill radius.
(ii) Recalculate miso at each time step to account for the
drop in Σg as the material in the planet's feeding zone di-
minishes.
(iii) Allow the planet to grow rapidly to miso by removing
gas from the disc around the planet and adding it to the
planet, using gas accretion rates obtained from the fits to
the Movshovitz et al. (2010) models. Once the planet reaches
miso, it transitions to type II migration and begins accreting
at the viscous rate.
dt ≥ 2 M⊕ per 1000 yr. When the gas isola-
When implementing the above prescription, we define
the point at which the planet enters runaway gas accretion
to be when dm
tion mass is calculated we assume a maximum gas isolation
mass of 400 M⊕, which accounts for when a planet enters the
runaway gas accretion phase in a massive disc, where tidal
torques from the planet would evacuate the feeding zone
before the gas isolation mass was reached. We note that a
planet that does not reach the runaway gas accretion mass
prior to reaching the local gap forming mass would instead
transition directly to type II migration without accreting the
material within its feeding zone, and will begin accreting at
the smaller of the rate obtained from the fits to Movshovitz
et al. (2010), or the viscous supply rate.
2.2.2 Migration during runaway gas accretion
Until the planet reaches the mass required for runaway gas
accretion, it undergoes type I migration using the torque
formulae of Paardekooper et al. (2010, 2011). Once it un-
dergoes runaway gas accretion, the planet begins to carve
a gap in the disc by rapidly accreting the surrounding ma-
terial. To account for this change in conditions, the planet
stops undergoing type I migration and begins to migrate at
a rate with a timescale equal to the local viscous evolution
time:
τν =
2rp
3ν
.
(3)
Migration at this rate continues until the planet reaches the
gas isolation mass, where it transitions to self-consistent type
II migration driven by the coupling to the viscous evolution
of the disc via the impulse approximation (Lin & Papaloizou
1986). We note that recent hydrodynamic simulations have
indicated that the migration of gap forming planets does
not necessarily occur at exactly the viscous flow rate of
the gas in the disc (Duffell et al. 2014; Durmann & Kley
2015), due to residual gas in the gap adding to the migration
torque. For the disc and planet masses that we consider in
4 G. A. L. Coleman and R. P. Nelson
this work, however, the migration rates provided by the im-
pulse approximation are in reasonable agreement with those
obtained in hydrodynamic calculations (Coleman & Nelson
2014).
2.3 Disc radial structures
Our recent simulations failed to form any surviving gas gi-
ant planets, other than hot Jupiters that are only prevented
from migrating into their host stars by the presence of a cen-
tral magnetospheric cavity (Coleman & Nelson 2014, 2016).
An analysis of the conditions required for gas giants to form
and survive outside of the central cavity presented in Cole-
man & Nelson (2014) demonstrated that runaway gas ac-
cretion and the transition to type II migration needs to oc-
cur when the planets are distant from their stars. For ex-
ample, for a Jovian mass planet to form and settle into a
final orbit at ∼ 1 au, requires type II migration to be ini-
tiated at ∼ 6 au. A Jovian planet orbiting at ∼ 5 au needs
to initiate runaway gas accretion and type II migration at
∼ 15 au. The time of formation also provides a constraint:
form too early in the disc life time and a planet migrates all
the way into the central cavity; form too late and there is
insufficient gas available to build a gas giant. It is notewor-
thy that population synthesis simulations produce a large
number of surviving cold gas giants (e.g. Mordasini et al.
2009). Coleman & Nelson (2014) examined the planet mass
and orbital evolution obtained using the following three ap-
proaches: 1D disc models similar to those presented in this
paper; 2D hydrodynamic simulations that were designed to
match the conditions in the 1D models; the prescriptions
for mass growth and migration used in population synthesis
models. They showed that the discrepancy obtained in gi-
ant planet survival rates between the modelling approaches
arises because a migration-slowing factor is included in the
population synthesis models when in the so-called planet-
dominated regime, and this results in too much slowing of
type II migration compared to that observed in the 2D hy-
drodynamic simulations or in the 1D viscous disc models.
Retaining the cores of gas giants at large orbital radii is
difficult, especially late in the disc life time. The corotation
and Lindblad torques need to balance, such that the core or-
bits in a "zero-migration zone" (Bitsch & Kley 2011; Hellary
& Nelson 2012; Cossou et al. 2013). The corotation torque
has entropy-related and vortensity-related components, and
it is the entropy-part that is normally strongest and able
to balance the Lindblad torque when the temperature pro-
file decreases outwards steeply. In a viscous, irradiated disc,
the inner regions of the disc, where the viscous dissipation
dominates the heating, have steep temperature gradients,
and early in the disc life time the zero-migration zone can
extend out to ∼ 10 au for planet masses (cid:38) 10 M⊕ (Bitsch
& Kley 2011; Hellary & Nelson 2012; Cossou et al. 2013;
Coleman & Nelson 2014; Bitsch et al. 2015a). As the disc
evolves, however, the viscous heating rate decreases and the
zero-migration zone moves into the inner 1–2 au and only
prevents the migration of lower mass planets. Although the
details of the evolution depend on input parameters such as
the viscous stress and the opacity, it would seem to be dif-
ficult to maintain strong entropy-related corotation torques
in the outer disc regions during the later phases of disc evo-
lution. One alternative for maintaining cores at large radii
might be for the vortensity-part of the corotation torque
to be strengthened in regions where the surface density in-
creases with radius, such as may occur if the disc surface
density contains undulations . These regions might act as
planet traps (Masset et al. 2006), as well as being regions
where small sized bodies such as dust, pebbles, boulders
and small planetesimals could concentrate (e.g. Pinilla et al.
2012). The main focus of this paper is to examine the conse-
quences of allowing protoplanetary discs to be radially struc-
tured because of radial variations in the viscous stress. Our
approach is to employ a very simple "toy model" for simulat-
ing these radial structures, but we derive motivation from re-
cent observations of protoplanetary discs, and from the long
history of MHD simulations showing that discs which sup-
port magnetorotational turbulence (Balbus & Hawley 1991)
often demonstrate radial structuring in the form of zonal
flows.
2.3.1 Observed structures
Recent observations of the young class I T Tauri star,
HL Tau, have shown the presence of a number of quasi-
axisymmetric rings, corresponding to maxima and minima
in the emitted intensity as a function of radius. The sys-
tem of rings extends between 13–100 au (ALMA Partnership
et al. 2015). A number of suggestions have been put forward
to explain the rings, included embedded planets (Dipierro
et al. 2015; Picogna & Kley 2015), pressure bumps that
trap dust (Flock et al. 2015), enhanced dust growth near ice
lines (Zhang et al. 2015), and sintering of dust aggregates
(Okuzumi et al. 2015). Even more recent ALMA observa-
tions of the disc around TW Hydra have also uncovered a
series of rings (Andrews et al. 2016), suggesting that these
really are common phenomena that arise during the evolu-
tion of protoplanetary discs. The closer proximity of TW
Hydra to the Solar System allows regions of the disc that
lie closer to the central star to be probed by the ALMA
observations, and these have uncovered rings between or-
bital radii 1 – 40 au. Furthermore, although high resolution
ALMA images of other protoplanetary discs have not yet
been released, existing ALMA data for a number of other
discs indicate that ring structures are present in the disc
outer regions (Zhang et al. 2016), suggesting that these fea-
tures are common phenomena that arise during the evolution
of protoplanetary discs. Although we do not attempt to fit
our simulations to these observations, we simply note that a
plausible scenario for the origin of these rings is radial varia-
tion in the effective viscous/turbulent stresses that give rise
to variations in the surface density.
2.3.2 Zonal flows in MHD simulations
Over a number of years, both global (Steinacker & Pa-
paloizou 2002; Papaloizou & Nelson 2003; Fromang & Nelson
2006) and local (Johansen et al. 2009) simulations of magne-
tised discs have demonstrated the occurrence of persistent
density/pressure maxima and minima as a function of ra-
dius, arising from localised magnetic flux concentration and
associated enhancement of magnetic stresses. More recent
simulations incorporating non-ideal MHD effects have also
reported the existence of these features in local (Bai & Stone
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
5
stresses may be a common feature of planet forming discs. In
addition to the zonal flows described above, similar features
may also arise in regions where there is a transition from one
non-ideal MHD process being dominant to another becom-
ing dominant (e.g. a transition between Hall and ambipolar
dominated regimes), or at the interface between magneti-
cally active and dead zones. In the presence of these transi-
tions, the disc may not be able to maintain a constant mass
flux through all radii at all times, and radial structuring may
occur. For simplicity, in this paper we just consider a rather
crude model for disc structuring that is intended to mimic
the growth and decay of zonal flows, but we note that radial
structuring may also occur because of other physical pro-
cesses that influence the local rate of mass flow through the
disc.
2.3.3 A simple model for radial structuring
We incorporate radial structuring in our models by intro-
ducing a spatially and temporally varying viscous stress.
At any one time, four structures are present in our simu-
lations. While this number is arbitrary, it is similar to the
number of rings observed in HL Tau and TW Hydra. Each
one exists between specific, predefined radii (Rmin, Rmax),
where the values are given in Table 1. Each structure has
a finite life time (see the final column of Table 1), and as
it decays a new structure grows within the same range of
radii Rmin < R < Rmax. We initiate the structures 50,000
years after the start of the simulations, once the disc has
reached a quasi-steady state, by increasing the viscosity pa-
rameter α up to a maximum strength of 1.5× that of the
background value. This value was chosen to approximately
match the ∼ 50% variation in the surface density due to
the zonal flows obtained in the MHD simulations of Bai &
Stone (2014). For each structure, the maximum value of α
is located at the centre of that structure, whilst we transi-
tion α to its background value over a distance of 3.5 local
scale heights using a Gaussian kernel, giving each structure
a width of 7 local scale heights. Once each of the structures
begins to form, it does so over 100 local orbital periods by
increasing α from the background value up to the required
value, as described below:
α(r, t) = αb +
αb
2
Rnewtnew
(4)
where αb is the background value, and Rnew and tnew are
the radial and time factors defined by
(cid:19)
(cid:18)−(R − Rstruc)2
(cid:18)
2H 2
struc
(cid:18) 6(t − tstart − 0.5(t100 − tstart))
t100 − tstart
Rnew = exp
tnew = 0.5 ×
tanh
(cid:19)
(5)
(cid:19)
+ 1
(6)
where the subscript 'struc' denotes the radial location of
the centre of the structure, tstart is the structure formation
time, t100 − tstart represents the time interval of 100 orbital
periods after the structure begins to form, evaluated at the
structure's centre, and H is the local disc scale height. The
shape and time evolution of the locally varying viscous α
parameter associated with an individual structure as it forms
Figure 1. Plot showing the time variation of the viscous α asso-
ciated with the formation of a radial structure over a time of 100
local orbits.
Structure Rmin Rmax
(au)
label
(au)
Lifetimes
(×103 local orbits)
1
2
3
4
4.25
9.25
14.25
19.25
5.75
10.75
15.75
20.75
10, 50, 100
10, 50, 100
10, 50, 100
10, 50, 100
Table 1. Radial structure parameters
2014) and global (Zhu et al. 2014; B´ethune et al. 2016) sim-
ulations. Density variations with amplitudes up to ∼ 50%
of the background have been reported (Bai & Stone 2014).
Being in geostrophic balance, these pressure bumps are of-
ten referred to as zonal flows (Johansen et al. 2009). Simon
et al. (2012) have recently observed long-lived zonal flows in
simulations with radial domains up to 16 scale heights, and
in these large shearing boxes they find that the outer radial
scale of the zonal flows is ∼ 6H, although they stress that
simulations in larger domains are required to demonstrate
convergence. Dittrich et al. (2013) ran shearing box simu-
lations with radial domains up to 21H and also found the
radial sizes of the axisymmetric zonal flows to be between
5 and 7H. The study by (Bai & Stone 2014) noted that
zonal flows in radially-narrow shearing boxes tended to be
intermittent, but runs in large shearing boxes of width 16H
persisted for the full duration of the simulations, which had
total run times of 400 orbits. Zonal flows are clearly able to
live for long times, but at present it is not clear what their
characteristic life times are.
Although we do not try to fit a model to these MHD
simulations, and instead take the approach of employing
a simple prescription to demonstrate "proof of concept",
we note that global MHD simulations which display dust
concentration in pressure bumps have been used to com-
pare theoretical calculations with the observed structures in
protoplanetary discs (Flock et al. 2015). Density and pres-
sure bumps arising from variations in magnetic or turbulent
MNRAS 000, 000–000 (0000)
6 G. A. L. Coleman and R. P. Nelson
Figure 2. Surface density profiles at t = 0.1, 1, 2, 3 Myr for a 1 × MMSN disc (total lifetime ∼ 5.5 Myr) without (left panel) and with
(right panel) radial structuring.
is shown in Fig. 1, where the α parameter gradually increases
to the required value, while maintaining a smooth profile.
Our structures have specific lifetimes, and this is a pa-
rameter that we vary in the simulations described below.
When a structure comes to the end of its lifetime it quickly
disappears over 100 local orbital periods. As one structure
disappears, another one forms at a randomly chosen location
within the range of allowed radii given in Table 1. When the
structure starts to disappear, α evolves according to:
α(r) = αb +
Rold(1 − told),
where told is given by
told = 0.5 ×
tanh
(cid:18) 6(t − tend − 0.5(t100 − tend))
t100 − tend
αb
2
(cid:18)
(cid:19)
(7)
(cid:19)
+ 1
.
(8)
Here, tend is the time at which the structure begins to dissi-
pate and t100 represents 100 orbital periods after this time.
Rold is equal to equation 5 but with values taken for the old
structure instead of a new one, as shown by this expression
(cid:18)−(R − Roldstruc)2
(cid:19)
2H 2
oldstruc
Rold = exp
.
(9)
To account for a new structure being influenced by a dissi-
pating older structure, equation 4 becomes
(Rnewtnew + Rold(1 − told))
α(r, t) = αb +
(10)
αb
2
This allows a smooth transition between two adjacent form-
ing/dissipating structures.
Below we discuss the main effects of radial structures on
the disc profile and migration of embedded low mass planets.
2.3.4 Effects on disc and planet evolution
Fig. 2 shows the surface density evolution for a 1 × MMSN
disc in simulations without (left panel) and with (right
panel) radial structuring. The drop in surface density in the
inner regions of the discs arises because both models include
an increase in α by a factor of 5 from the background values
of either 2 × 10−3 or 6 × 10−3 where the disc temperature
exceeds 1000 K (Coleman & Nelson 2016). The presence of
the radial structures arising from the variations in α in the
outer disc are evident in the right panel. While these surface
density dips have little influence on the global disc evolution,
they have a dramatic effect on planet migration.
Fig. 3 shows contours that illustrate the direction and
speed of type I planet migration as a function of planet
mass and semimajor axis at different times in a 1 × MMSN
disc with solar metallicity. The left panel shows a simula-
tion without radial structuring, and the right panel is for a
run with structuring switched on. Red regions correspond
to rapid inward migration, blue regions correspond to rapid
outward migration, and white contours interspersed between
the red and blue contours represent zero-migration zones
where Lindblad and corotation torques cancel. The white
contours at the top of the panels correspond to planets reach-
ing the gap opening mass and undergoing type II migration.
The planet trap arising from the inner fully-developed tur-
bulent region is represented by the innermost blue contour,
apparent in the first three frames of each simulation. As the
surface density decreases, the zero-migration zones and ex-
tended regions of outward migration associated with strong
entropy-related corotation torques slowly move in towards
the star on the disc evolution time scale in both runs, but
the run with radial structures maintains four zero-migration
zones in the outer disc for the duration of the simulation,
leading to the possibility of long-term trapping of planetary
cores with masses up to ∼ 30 M⊕.
If a planet core was to migrate to the edge of one of
the four structures, then it would be trapped for the life-
time of the structure. The core is released from the structure
when it comes to the end of its life, and the planet starts
to migrate inwards. A new structure is formed locally to re-
place the old one, and this has some probability of being
located inside the old one (that depends on the location of
the old structure within its allowed range of radii). If the
new structure sits inside the old one then the planet core
can be trapped by it, but if it sits outside the planet lo-
cation then the planet migrates inwards, either into one of
the other three structures, or in towards the star if it has
just escaped from the innermost structure. Furthermore, a
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
7
Figure 3. Contour plots showing regions of inwards (red) and outwards (blue) migration) in a 1 × MMSN disc at t = 0.1 (top left), 1
(top right), 2 (bottom left) and 3 Myr (bottom right) for discs without (left panel) and with radial structuring (right panel). The white
contours at the top of each panel corresponds to the planet reaching the local gap forming mass, at which point the planet will undergo
type II migration. The contours represent values of γΓ/Γ0, where γ is the ratio of specific heats, Γ is the torque experienced by a planet
and Γ0 is a normalisation factor defined in Paardekooper et al. (2010)
rapidly migrating planet core can escape from a structure
while it is decaying and before the next structure has devel-
oped fully. This shows that the long term orbital evolution of
a planetary core has a stochastic element that depends on
the detailed evolutionary histories of the radial structures
in the disc. Some cores remain trapped at large radius over
the disc lifetime, whereas other cores escape from the planet
traps and migrate into the disc inner regions.
Gas accretion can occur onto a core that is trapped if
its mass exceeds mp ≥ 3 M⊕, and if it remains in the outer
disc for an extended period of time then runaway gas ac-
cretion can occur and a giant planet can form. The planet
would then open a gap in the disc, and begin to undergo
type II migration as the planet traps are not effective for
gap forming planets. The process of building planets at the
planet traps is enhanced by the concentration of boulders
and planetesimals at these locations, which can then be ac-
creted efficiently by the growing planets. In general, we find
that accretion of solids by planetary embryos occurs during
an early burst, prior to the onset of the main gas accretion
phase. Gas accretion is then accompanied by modest plan-
etesimal accretion at rates that are similar to or below those
prescribed in the Movshovitz et al. (2010) models that deter-
mine gas accretion rates in our simulations. Approximately
20% of the giant planets in our runs experience an episode of
rapid and short-lived solids accretion, normally during the
runaway gas accretion phase when the growth of the giant
acts to destabilise the system. This burst of accretion can
either arise from an impact with a low mass protoplanet,
or through accretion of a local swarm of planetesimals over
a time period that is less than ∼10,000 years. Our fits to
the Movshovitz et al. (2010) models do not allow the gas
accretion rate to respond to this time-varying planetesimal
accretion, and this is one area for future improvement of our
model.
In summary, we have introduced a simple model for
the radial structuring of protoplanetary discs that includes
assumptions about the number of surface density features
MNRAS 000, 000–000 (0000)
(planet traps) that are formed and their lifetimes. We
present this model as a simple proof-of-concept in this paper,
and do not include an extensive analysis of what happens
when the model parameters are modified. It is reasonable
to suppose, however, that reducing the number of planet
traps and their lifetimes will result in less efficient trapping
of planet cores, and hence less efficacious giant planet forma-
tion. Precisely how the formation of giant planets is affected
by variation of model parameters will be examined in future
work.
2.4 Initial conditions
Table 2 gives an overview of the parameters used in our simu-
lations. All simulations were initiated with 44 planetary em-
bryos, of mass 0.2 M⊕, with semimajor axes between 1 and
20 au and separated by 10 mutual Hill radii. These were em-
bedded in a swarm of thousands of planetesimals/boulders,
that were distributed with semimajor axes between 0.5 and
25 au, with masses either 10, 20 or 50 times smaller than
the embryos, depending on the metallicity of the system.
(This varying mass ratio between embryos and planetesi-
mals was implemented to obtain a planetesimal number that
allowed the simulations to run on reasonable time scales. Be-
tween 1000 and 5000 planetesimals/boulders were used and
run times for individual simulations varied between 2 and
6 months.) The total mass of solids ranges between 12.5–
109 M⊕ depending on the disc mass and metallicity. The
effective physical radii of planetesimals were set to 10 m,
100 m, 1 km or 10 km, such that the primary feedstock of
the accreting protoplanets ranged from being boulders to
large planetesimals whose evolution differed principally be-
cause of the strengths of the gas drag forces that they experi-
enced. Initial eccentricities and inclinations for protoplanets
and planetesimals/boulders were randomized according to a
Rayleigh distribution, with scale parameters e0 = 0.01 and
i0 = 0.25◦, respectively.
8 G. A. L. Coleman and R. P. Nelson
Parameter
Disc mass
Disc metallicity
Total solids mass
Background viscous α
Planetesimal radii
Planetesimal mass
Planetesimal number
Gas disc lifetimes
Values/Ranges
1, 2 ×MMSN
0.5, 1, 2 × Solar
12.5–109 M⊕
2 × 10−3, 6 × 10−3
10 m, 100 m, 1 km, 10 km
0.004, 0.01, 0.02 M⊕
1000 – 5000
3.5 – 8.4 Myr
Table 2. Values, and the ranges of values, adopted for various
simulation parameters.
Collisions between protoplanets and other protoplanets
or planetesimals resulted in perfect sticking, which probably
results in a slight overestimate of accretion rates in the sim-
ulations. We neglect planetesimal-planetesimal interactions
and collisions in our simulations for reasons of computational
speed.
The gas disc masses simulated were 1 and 2 times the
mass of the minimum mass solar nebula (MMSN Hayashi
1981). We also vary the metallicity so that the initial solids-
to-gas mass ratios in the discs are equal to 0.5, 1 and 2 times
the solar value for the different models. We define the solar
metallicity to be equivalent to the solids-to-gas ratio intro-
duced by Hayashi (1981). We smoothly increase the mass of
solids exterior to the snow line by a factor of 4 by increasing
the numbers of planetesimals, and the initial surface den-
sity of solids follows the initial gas surface density power
law, as described in Hellary & Nelson (2012). The factor
of four increase in solids implies a larger increase in water
content than obtained by Lodders (2003) for the solar neb-
ula, but is consistent with the widely used Hayashi model
(Hayashi 1981). We adopt this model to maintain continuity
with our earlier work (Hellary & Nelson 2012; Coleman &
Nelson 2014, 2016). We track the changes in planetary com-
positions throughout the simulations, as planets can accrete
material originating either interior or exterior to the snow
line.
We use two different values for the background α value,
α = 2 × 10−3 and 6 × 10−3. These values of α correspond
to disc lifetimes of 5.5 and 3.5 Myr respectively for a disc
with mass equal to 1 × MMSN. We examine the effect of
varying the lifetimes of the radial structures in the disc, with
the three values assumed being 104, 5 × 104 and 105 local
orbital periods. We ran two instances of each parameter set,
where only the random number seed used to generate initial
particle positions and velocities was changed, meaning that
a total of 288 simulations have been run. The simulations
were run for 10 Myr, or until no protoplanets remained.
2.5 Planet classification scheme
To assist in the discussion of simulation outcomes, we have
developed a classification system for the different bodies that
are formed. As there are no formal IAU definitions for ex-
oplanet classes relating to their masses and compositions,
there is freedom of choice in how planets should be classi-
fied. We have chosen a scheme that uses mass as the pri-
mary discriminant and composition as a secondary one. We
use the labels "Earth", "Neptune" and "Jupiter", along with
the prefix "super" to define six mass-based classes, and sub-
classes are defined according to the volatiles content, either
in the form of ice or gas, of the planets. Definitions of the
different planet classes are given in Table 3. Note that when
we use the term "gas giant" we are referring to Jupiters or
super-Jupiters.
3 RESULTS
We now present the results for our simulations. We begin
by discussing a representative run in which multiple giant
planets were able to form and survive. We then present an
overview of all the simulation outcomes, before examining
how modifying parameters such as the disc mass, metallicity,
photoevaporation model etc. changes the results.
3.1 Run CJ120.1210A
Run CJ120.1210A had a disc mass of 1 × MMSN, 2× solar
metallicity, and contained planetesimals with radii Rp =
100 m. The total mass in planetesimals was 43.2 M⊕ and
that in protoplanets was 8.8 M⊕. The background α = 2 ×
10−3, and radial structures had a lifetime of 10,000 local
orbits. The direct photoevaporation model was used.
The evolution of protoplanet masses, semimajor axes
and eccentricities are shown in Fig. 4, and the final state of
the system is also represented in Fig. 10 (the case with the
label CJ120.1210A). We also show the mass versus orbital
period evolution of all protoplanets in Fig. 5, where filled
black circles represent surviving planets, and the evolution
of the labelled planets is described below. The end state af-
ter 10 Myr consists of: an inner compact system comprising
3 super-Earths/Neptunes; a cool Neptune and Earth-mass
planet orbiting between 2.5–3.4 au; two cold Jupiters orbit-
ing between 6–12 au; a collection of low mass planets ('de-
bris'), that failed to grow during the simulation, orbiting out
beyond 20-30 au. We ignore the long period 'debris' in our
discussion below, and just concentrate on the other planets
that form.
3.1.1 Cold Jupiters
The cores of the two Jupiters (see planet labels 1 and 2 in
Figs. 4 and 5) begin to form at orbital radii 15-25 au within
the first 0.5 Myr of the simulation, through a combination
of planetesimal accretion and mutual collisions between em-
bryos. Migration and trapping of planetesimals in the radial
structures helps concentrate material which is then accreted
by the embryos, stimulating rapid growth above 3 M⊕ such
that gas accretion onto the growing cores can start. These
proto-giant planets remain trapped at large radii by the ra-
dial structures, and continue to accrete gas steadily until
runaway gas accretion is initiated at times just before and
after 3 Myr, respectively (see the top panel of Fig. 4). The
rapid burst of gas accretion takes the planet masses up to
∼ 100 M⊕, after which gap opening ensues. Initially both
planets accrete at the viscous supply rate, but 'planet 1'
truncates the disc exterior to it and prevents further gas
accretion on to 'planet 2', which lies interior to 'planet 1'.
The onset of gap formation allows the planets to type II
migrate inwards until the gas disc is completely removed
MNRAS 000, 000–000 (0000)
Classification
Mass
Rock % Ice % Gas % Final Number
Giant planet formation in structured discs
9
Rocky Earth
Water-rich Earth
Rocky super-Earth
Water-rich super-Earth
Gas-rich super-Earth
Gas-poor Neptune
Gas-rich Neptune
Gas-poor super-Neptune
Gas-rich super-Neptune
Jupiter
Super-Jupiter
mp < 3 M⊕
mp < 3 M⊕
3 M⊕ ≤ mp < 10 M⊕
3 M⊕ ≤ mp < 10 M⊕
3 M⊕ ≤ mp < 10 M⊕
10 M⊕ ≤ mp < 35 M⊕
10 M⊕ ≤ mp < 35 M⊕
35 M⊕ ≤ mp < 100 M⊕
35 M⊕ ≤ mp < 100 M⊕
100 M⊕ ≤ mp < 1000 M⊕
mp ≥ 1000 M⊕
> 70
< 70
> 60
N/A
N/A
N/A
N/A
N/A
N/A
N/A
N/A
< 30
> 30
< 30
> 30
N/A
N/A
N/A
N/A
N/A
N/A
N/A
0
0
< 10
< 10
> 10
< 10
> 10
< 50
> 50
> 50
> 50
1563
4625
12
83
69
5
79
29
147
120
12
Table 3. Planetary classification parameters based on their mass and composition. Note that water-rich planets are so-called because
they accrete water ice in solid form that originates from beyond the snow-line. Characteristics that play no role in the classification of
a planet are denoted by "N/A" in the relevant columns. Note all Jupiters and Super-Jupiters formed in the simulations had gas mass
fractions ≥ 50%.
Figure 4. Evolution of masses, semimajor axes and eccentricities of all protoplanets in simulation CJ120.1210A, where the disc lifetime
∼ 4.5 Myr. Note that formation histories of selected surviving planets are indicated by the labels on the right side of the mass and
semi-major axis subplots.
after ∼ 4.5 Myr, although we note that the migration of
'planet 2' is slowed by the truncation of the disc by 'planet
1'. The gas giants have masses 306 M⊕ and 222 M⊕, gas mass
fractions of 98%, semimajor axes 6.3 au and 11.4 au, orbital
periods 15.8 yr and 38.5 yr and eccentricities ∼ 0, respec-
tively. While this pair of planets are far from being a perfect
analogue to the Jupiter-Saturn system, there is an obvious
MNRAS 000, 000–000 (0000)
similarity in terms of gross characteristics that is worthy of
note (i.e. mass of inner planet > mass of outer planet and
semimajor axis ratio ∼ 1.8).
10 G. A. L. Coleman and R. P. Nelson
Figure 5. Evolution of planet mass versus orbital period for all
protoplanets in simulation CJ120.1210A. Filled black circles rep-
resent final masses and orbital periods for surviving planets. Note
that formation histories of selected surviving planets are indicated
by the labels adjacent to the filled black circles. The dotted black
line at ∼ 4d represents the inner edge of our simulated proto-
planetary disc. The arrows above the x-axis indicate the average
positions of the four radial structures.
3.1.2 Cool Neptune and Earth
These planets are labelled as 3 and 4 in Figs. 4 and 5. The
cool Neptune begins its formation out beyond 10 au at the
same time as the giant planet cores are forming, but interior
to these two proto-giants. It also begins to accrete gas within
the first 0.5 Myr, but at a slightly slower rate than the two
proto-giants, and remains trapped by the radial structures
during the first 3 Myr. The cool Neptune is nudged inwards
when the innermost gas giant undergoes runaway gas ac-
cretion and starts type II migrating, and this allows the
Neptune to escape the radial structures and migrate in to-
wards the central star. Gas accretion onto the Neptune and
its migration halt when the gas disc disperses after 4.5 Myr,
leaving it with a mass of 28.6 M⊕, gas mass fraction of 86%,
semimajor axis ∼ 3.5 au, orbital period 6.5 yr, and eccentric-
ity ∼ 0. As this gas-rich Neptune escaped the radial struc-
tures and migrated inwards it shepherded a 1 M⊕ water-rich
terrestrial planet ahead of it, which had a final semimajor
axis ∼ 2.7 au and orbital period ∼ 4.4 yr.
3.1.3 Compact inner system of super-Earths/Neptunes
The planets we discuss here have labels 5-7 in Figs. 4 and
5. This compact system forms from a combination of bodies
that are initially orbiting interior to the radial structures
and one dominant body that originates from larger radii.
This more massive body grows through planetesimal accre-
tion and collisions with neighbouring embryos out beyond
10 au, where it starts to accrete gas and remains trapped
by the radial structures until 2 Myr. At this point its mass
is 5 M⊕, and it is able to escape from the radial structures
by migrating through them as they switch on and off, after
which it undergoes rapid inward type I migration while con-
tinuing to accrete gas (becoming a gas-rich Neptune in the
process). The gas-rich Neptune shepherds a large number of
interior embryos in a resonant convoy as it migrates, and
when the gas disc starts to disperse after ∼ 3.5 Myr this
convoy breaks up and mutual collisions between the numer-
ous embryos lead eventually to the formation of a compact
inner system comprised of 3 planets: a gas-poor Neptune
with mass 11.6 M⊕, gas mass fraction 7%, semimajor axis
∼ 0.07 au, orbital period 6 days and eccentricity 0.11; an
icy super-Earth with mass 8.4 M⊕, gas mass fraction 6.5%,
semimajor axis ∼ 0.15 au, orbital period 16.2 days and ec-
centricity 0.2; a gas-rich Neptune with mass 27.6 M⊕, gas
mass fraction 56%, semimajor axis ∼ 0.2 au, orbital period
33.2 days and eccentricity 0.07. We note that the eccentrici-
ties of the these planets were pumped up to the values shown
in the bottom panel of Fig. 4 during a late scattering event
at 5.2 Myr.
3.2 Ensemble results
We now discuss the results of the simulations as a whole,
focusing first on the masses and periods of the planets that
form, and then on the eccentricity distribution.
3.2.1 Masses and periods
Considering the results of the simulations as a whole, 132
surviving giant planets are formed with masses ranging from
0.3 MJupiter to 4 MJupiter, with periods from 5 days up to
24000 days (the smaller period being determined by our
boundary conditions). The majority of these giant planets
formed at the outer edges of radial structures, whilst a hand-
ful of less massive giant planets accreted the majority of their
gas envelopes after escaping from the radial structures and
type I migrating towards the central star. Fig. 6 shows a
mass versus period diagram for all of the surviving planets
from the simulations, along with all confirmed exoplanets
(Han et al. 2014).
The known exoplanets form three apparently distinct
groups in the mass-period diagram: cold Jupiters with or-
bital periods (cid:38) 100 days; hot Jupiters with orbital periods
(cid:46) 10 days; super-Earths/Neptunes with periods between
2 (cid:46) P (cid:46) 100 days. These features are affected by a number
of observational biases, including the fact that ground based
transit surveys are only sensitive to Jupiters with periods
(cid:46) 10 days. Nonetheless, analysis of the period distribution
of planets detected only by radial velocities seems to con-
firm that there is a real valley in the distribution between
10–100 days (Udry et al. 2003; Cumming et al. 2008). More
recently, Santerne et al. (2015) have presented an analysis
of giant planets discovered by the Kepler spacecraft that
were followed-up using radial velocity measurements over 6
years, and they confirm that the period-valley also exists
within this data set. One of the most striking features when
comparing the results of the simulations with the observa-
tional data in Fig. 6 is the fact that the giant planets formed
in the simulations are almost all hot Jupiters (periods < 10
days) and cold Jupiters (periods > 100 days), with only a
few massive bodies being located in the region that corre-
sponds to the observed period valley. Furthermore, the sim-
ulations produce numerous planets with masses in the range
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
11
Figure 6. Left panel: Mass vs period plot, comparing observed exoplanets (red squares) with our simulation results (blue circles) and
the Solar System planets (black diamonds). Right panel: Same as left panel but data taken from Coleman & Nelson (2016). The grey
zones indicate the habitable zone (Kasting et al. 1993). The arrows at the bottom of the left panel indicate the average positions of the
four radial structures.
0.5 M⊕ (cid:46) mp (cid:46) 30 M⊕ and periods between 2 (cid:46) P (cid:46) 100
days, that correspond to the observed super-Earths and Nep-
tunes. Some of these lower mass planets are in systems that
contain giant planets, as described in the previous section
for run CJ120.1210A, and some are devoid of any giants.
Comparing the left panel of Fig. 6 with the right panel (a
reproduction of Fig. 12 in Coleman & Nelson (2016)), where
similar N-body simulations, but without the inclusion of disc
radial structures, were presented, we see that the agreement
between the observed and simulated planet distributions is
much improved in this work.
In the simulations, the origin of the two distinct popula-
tions of hot and cold Jupiters, and the period valley between
them, can be explained as follows. Giant planets that form
early in the disc lifetime migrate all the way into the mag-
netospheric cavity, and become hot Jupiters. Giant planets
that are destined to not become hot Jupiters must form
near the end of the disc lifetime, when photoevaporation
plays an important role in the disc evolution. Photoevap-
oration, combined with viscous evolution, causes the disc
to disperse from the inside out. There is a high probability
that a giant forming towards the end of the disc lifetime
will migrate towards the star when the disc interior to the
critical radius for photoevaporation has been fully or par-
tially evacuated, preventing it from migrating close to the
star, and ensuring that it remains as a cold Jupiter. Hence,
the observed giant planet period distribution may arise as
a combination of forming giant planets at large radius, hav-
ing a stopping mechanism for migration at the inner edge
of the disc (i.e. a magnetospheric cavity) and the inclusion
of photoevaporation, which occurs outside a well-defined ra-
dius corresponding roughly to where a thermal wind can be
launched. The influence of different models of photoevapo-
ration on the results are discussed in more detail below, but
we note that Alexander & Pascucci (2012) have suggested
that disc clearing due to photoevaporation can be responsi-
ble for a pile-up of giant planets at 1 au, as planet migration
is slowed when photoevaporation begins to dominate disc
evolution. More recently Ercolano & Rosotti (2015) showed
that different models of photoevaporation influence the pile-
MNRAS 000, 000–000 (0000)
up location, with a thermal-wind launching inner radius of
1–2 au being preferred.
Low mass, compact systems that formed and migrated
to the inner regions of the disc are seen in a number of
simulations. The formation of these compact systems occurs
similarly to those described in Coleman & Nelson (2016),
but some compact systems within this work contained giant
planets with large orbital periods, as shown in sect. 3.1. The
co-existence of long period giant planets and low mass com-
pact systems in the simulation results seems to be in accord
with the recent analysis of Kepler data indicating the pres-
ence of long period giant planets around stars known to host
compact multi-systems (Uehara et al. 2016; Kipping et al.
2016).
3.2.2 Eccentricities of giant planets
The eccentricity distribution of observed giant (mp sin i ≥
100 M⊕) exoplanets is shown in Fig. 7 for bodies with orbital
periods > 10 days, along with the eccentricity distribution
for planets in the same mass and period range that form
in the simulations. It is clear that the eccentricity distribu-
tion associated with observed exoplanets is much broader
than that generated in the simulations. The maximum ec-
centricity for a giant planet obtained in the simulations was
ep = 0.13, whereas significant numbers of exoplanets are
observed to have eccentricities > 0.3. We note that those
simulated systems that resulted in modestly eccentric giants
did so because the giant planets underwent strong gravita-
tional scattering with other planets in the system, where the
scattered bodies typically had masses (cid:39) 20 M⊕. Scattering
between more massive bodies is required to obtain the larger
eccentricities observed in the exoplanet data (e.g. Rasio &
Ford 1996).
Given that our simulations end after 10 Myr, it is pos-
sible that dynamical instabilities could occur on longer time
scales in systems containing multiple giant planets, changing
the statistics shown in Fig. 7. We have examined the distri-
bution of mutual semimajor axis separations, expressed as
a function of mutual Hill radii, to determine whether or not
12 G. A. L. Coleman and R. P. Nelson
Figure 7. Distribution of observed giant exoplanet eccentricities
(blue) and the distribution arising from the simulations (red).
this is possible. We note that Marzari & Weidenschilling
(2002) examined the dynamical stability of three Jovian-
mass planets on initially circular orbits, and demonstrated
that the instability time scale for such a system scales with
the mutual Hill radius separation, with systems separated by
∼ 6 mutual Hill radii having instability times of ∼ 109 yr. All
of our systems are at least as separated as this, with approx-
imately half of the systems having semimajor axis separa-
tions between 6 and 12 mutual Hill radii, and the other half
being more separated. This suggests that some of the simu-
lated systems may undergo dynamical instabilities on time
scales longer than 10 Myr, but it seems highly unlikely that
running the simulations for Gyr time scales would result in
an eccentricity distribution that matches the observed one.
Assuming that the observed eccentricity distribution of
giant exoplanets arises primarily because of dynamical in-
stabilities in multiplanet systems, and using the observed
distribution as a constraint on viable formation scenarios,
the data suggest that giant planets must often form in com-
pact configurations, and do so more frequently than occurs
in our simulations.
Finally, we note that our simulations adopt a highly
simplified prescription for the eccentricity damping experi-
enced by gap forming planets, namely that the eccentricity is
damped on a time scale of ∼ 10 planet orbits. This is applied
independently of the mass remaining in the gas disc, and so
acts to bias our final systems towards having low eccentrici-
ties by reducing the likelihood of instabilities occurring while
the gas disc is present. It is clear that a more sophisticated
model will need to be adopted in future simulations if a more
realistic assessment of the ability of the models to generate
high eccentricity systems is to be undertaken.
3.3 Different photoevaporation models
3.3.1 Direct photoevaporation
Simulation CJ120.1210A, presented in sect. 3.1, was one of
a group of simulations that allowed direct photoevaporation
Figure 8. Normalised cumulative distribution functions of giant
planet periods for radial velocity (black dashed line) and Kepler
observed giant planets (black dot-dash line), and simulations with
different photoevaporation regimes; direct (blue line), standard
(green line), and none (red line). We define a giant planet in both
simulations and observations as a planet with mass mp ≥ 100 M⊕.
to impact the disc when the gas disc interior to the critical
radius had accreted onto the central star. This can occur
when a gap forming planet forms exterior to the critical
photoevaporation radius, and the inner disc drains onto the
star. In this scenario, the giant planet assists its own survival
against migration by stimulating the onset of direct photo-
evaporation and reducing the disc lifetime. Fig. 8 compares
the cumulative distributions of giant planet periods from
simulations with different photoevaporation models (colored
lines) and observations (black lines). When comparing the
observations, it is evident that for giant planets observed by
Kepler, the ratio of hot Jupiters to cold Jupiters is lower than
that found by radial velocity surveys. One possible reason
for this is that the average of the metallicities of the Kepler
stars is -0.18 dex (Huber et al. 2014), and this is lower than
for stars in the solar neighbourhood where the average is
-0.08 dex (Sousa et al. 2008). Comparing the observations
with our simulations, it is clear that the blue line, repre-
senting simulations with direct photoevaporation, compares
very reasonably with the observations, albeit with a higher
fraction of hot Jupiters. Given that the simulations shown
here have an average metallicity of 0.3 dex, the increased
ratio is perhaps unsurprising, given that the boost in solid
material can allow more rapid planet formation and there-
fore more time for migration. The period valley discussed
above is also evident here, as is the good agreement between
the simulated and observed cold Jupiter distributions.
Having observed the effect that direct photoevapora-
tion has on the survival of giant planets with long orbital
periods, we ran a further two sets of simulations with the
same parameters as described in sect. 2.4, but with differ-
ent photoevaporation models, the standard one (obtained by
just switching off direct photoevaporation) and no photoe-
vaporation, in order to examine their effects on giant planet
formation. The results of all simulations with disc mass 1×
MNRAS 000, 000–000 (0000)
MMSN, metallicity 2× solar and α = 2× 10−3 are shown by
the red (no photoevaporation) and green (standard) lines in
Fig. 8 and are discussed in the following sections.
3.3.2 Standard photoevaporation
Given that only a modest number of our simulations contain-
ing large (cid:38) 1 km planetesimals formed giant planets, we only
ran simulations with 10 m boulders or 100 m planetesimals
to examine the influence of switching off direct photoevapo-
ration and retaining the standard photoevaporation model.
We note that when comparing the results of simulations
that employed standard and direct photoevaporation mod-
els, evolution of the disc and planets are identical until the
time that direct photoevaporation is activated. This means
that the formation pathways of giant planets is similar, and
significant differences only arise for those cases where giant
planet formation and migration occurs near to the end of
the disc lifetime, when photoevaporation is strongly influ-
encing the disc evolution. Direct photoevaporation causes
the disc to be removed more rapidly, and so is more effec-
tive at stranding migrating planets at larger orbital radii.
Planets that form and migrate in discs with standard pho-
toevaporation are therefore more likely to form hot Jupiters,
as indicated in Fig. 8.
3.3.3 No photoevaporation
In this set of simulations, we neglect photoevaporation en-
tirely, such that the only processes that can deplete the gas
disc are viscous evolution and accretion onto planets, signif-
icantly increasing disc lifetimes and the time periods over
which migration can occur. We consider only models con-
taining 10 m boulders and 100 m planetesimals.
The early formation and evolution of giant planets is
similar to that seen in simulations with photoevaporation.
Once a giant planet forms, however, the lack of an effective
disc removal mechanism means that it will almost always
migrate all the way to becoming a hot Jupiter, as shown
by the red line in Fig. 8, where 95% of the giant planets
formed are hot Jupiters. The giant planets that remain as
cold Jupiters only did so because they formed late in the
disc lifetime, where they survived migration by accreting
the majority of the remaining gas disc. This ratio of hot
Jupiters to cold Jupiters is not consistent with observations,
and shows that recreating the observed distributions of giant
planets is extremely difficult without a mechanism for disc
dispersal.
3.4 Evolution as a function of model parameters
We now discuss the effects that varying the model param-
eters have on the formation and evolution of giant planets
in the simulations. Since these effects are consistent across
all photoevaporation models employed, we will only discuss
the simulations that include direct photoevaporation. Fig. 9
shows the cumulative distributions for simulated planets as
a function of the different parameters considered.
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
13
3.4.1 Disc mass and metallicity
The simulation results show a strong dependence on the ini-
tial mass and metallicity of the disc. Simulations with small
disc masses and sub-solar metallicities (e.g. 1× MMSN and
0.5× solar metallicity) are unable to form any giant planets,
due to the quantity of solid material in the disc being insuf-
ficient to form a massive planet core capable of accreting a
massive gas envelope during the disc life time. Increasing the
inventory of solids by increasing the total disc mass, or by
increasing the metallicity, leads to the formation of giants.
We see from Fig. 9 that the 1× MMSN, 2× solar metallicity
runs form moderate numbers of hot Jupiters, with 90% of
the giant planets having periods > 100 days. This is for the
following reasons: the planet cores form quite late in the disc
lifetime; the disc lifetime is shorter than for heavier discs;
the low disc mass leads to slower type I migration. Increas-
ing the disc mass and metallicity can be seen to dramatically
increase the numbers of hot Jupiters, as planet cores form
earlier, type I migration is faster and the disc lifetime is
longer. Models with disc mass 2× MMSN and metallicity
2× solar form numerous giant planets, and 80% of these are
hot Jupiters.
3.4.2 Planetesimal radii
The cumulative distributions for the giant planet orbital pe-
riods formed in simulations with different planetesimal radii
are shown in the top-right panel of Fig. 9. No giant plan-
ets formed in simulations where the planetesimal size was
10 km, in agreement with the very anaemic growth found in
Coleman & Nelson (2016) for models with 10 km planetes-
imals. Large planetesimals do not migrate very far through
the disc during its lifetime, and the relatively weak damping
means that their accretion rate onto planetary embryos re-
mains small because of their large velocity dispersion. Accre-
tion rates are slightly higher for 1 km planetesimals, leading
to 12 giant planets forming in these runs. Overall, only ∼ 5%
of all giant planets formed do so in simulations with 1 or
10 km sized planetesimals (half of all runs). When the plan-
etesimal radius is decreased to 100 m, or we consider 10 m
boulders, then giant planets form easily. Coleman & Nelson
(2016) found that planetary growth is efficient in the pres-
ence of small bodies that experience strong gas drag, since
they can migrate over large distances (helping growing em-
bryos to exceed their local isolation masses), and maintain a
relative modest velocity dispersion due to strong eccentricity
and inclination damping. Our inclusion of radial structures
allows small planetesimals and boulders to concentrate, and
growing embryos to avoid rapid inward migration. Hence,
the simulations form surviving giant planets with a broad
range of orbital periods. Similar numbers of giant planets
formed in simulations with 10 m and 100 m small bodies,
while their orbital period distribution (i.e. number of hot
Jupiters versus cold Jupiters) was also similar, as is shown
by the cumulative distributions in the top-right panel of Fig.
9.
3.4.3 α viscosity
The bottom-left panel of Fig. 9 shows that a lower viscosity
(i.e. mass accretion rate through the disc for a given disc
14 G. A. L. Coleman and R. P. Nelson
Figure 9. Normalised cumulative distributions of simulated giant planets as a function of different parameters. Top-left panel: Disc mass
and metallicity. Top-right panel: Planetesimal radii. Bottom-left panel: α parameter. Bottom-right panel: Radial structure lifetimes.
Bracketed values represent the number of giant planets in those cumulative distributions. We define a giant planet in both simulations
and observations as a planet with mass mp ≥ 100 M⊕.
mass) gives rise to a larger ratio of hot to cold Jupiters. This
is an effect of the shorter disc lifetimes associated with more
viscous discs, by approximately 2 Myr in our simulations.
A closely related effect is that the numbers of giant planets
that form in higher viscosity discs is lower than in lower
viscosity discs: 88 formed in the α = 6 × 10−3 runs versus
106 in the α = 2 × 10−3 simulations.
3.4.4 Radial structure lifetime
The cumulative distributions of orbital period for runs with
different assumed lifetimes for the disc radial structures are
shown in the bottom-right panel of Fig. 9. It is clear that
varying these lifetimes between 104 and 105 local orbit pe-
riods has very little influence on the results. We expect that
shorter lifetimes than those considered in the runs would
reduce the numbers of giant planets that form, since more
growing cores could escape from the outer disc regions and
migrate rapidly into the inner magnetospheric cavity before
becoming giants. It is also likely that the mass distribution
of the giants would be skewed towards lower masses, and the
ratio of hot to cold Jupiters would increase. By decreasing
the lifetimes of the radial structures to very short values we
would eventually converge towards the results presented in
Coleman & Nelson (2016), where all surviving giant planets
were hot Jupiters and had sub-Jovian masses.
3.5 Planetary system architectures
We find a diversity in the planetary system architectures
arising from the simulations. An ensemble of simulated
planetary systems displaying different architectures are
shown in Fig. 10, where the different architectures are
represented by different simulation label prefixes. Below we
describe the different architectures, and the general physical
conditions and modes of evolution associated with each of
them:
(i) Low-mass planetary systems – These form in simulations
where protoplanet growth rates are insufficient to form
giant planets. In some cases these are compact planetary
systems, with similar formation histories to those discussed
in Coleman & Nelson (2016). The systems with the prefix
'CS' (compact system) in Fig. 10 show the final configu-
rations from these runs, where the lack of massive planets
is evident along with their compactness. Generally, these
systems arose in metal-poor low-mass discs with small
planetesimals/boulders, or in more massive discs with large
planetesimals (i.e. Rpl ≥ 1 km).
(ii) Lonely hot Jupiters – Systems containing only hot
Jupiters formed in massive metal-rich discs. Typically
multiple giant planets form in the outer regions of the disc
and migrate to become hot Jupiters, where only the last hot
Jupiter survives. Often accompanying these hot Jupiters
are low-mass planets on long period orbits (Pp ≥ 100d), as
shown by systems with prefixes 'HJ' (hot Jupiters) in Fig.
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
15
Figure 10. Plot comparing different architectures arising from the simulations, with the Solar System and Kepler-167 included for
comparison. Orbital period is indicated on the x-axis and planet masses are indicated by the symbol size (radius scales with the square-
root of the planet mass). The symbol colours indicate the classification of each planet: red = Earths (mp < 3 M⊕); green = super-Earths
(3 M⊕ ≤ mp < 10 M⊕); blue = Neptunes (10 M⊕ ≤ mp < 35 M⊕); orange = super-Neptunes (35 M⊕ ≤ mp < 100 M⊕); black = Jupiters
and super-Jupiters (mp > 100 M⊕). See Table 3 for definitions of planet types.
MNRAS 000, 000–000 (0000)
16 G. A. L. Coleman and R. P. Nelson
10. From an observational perspective, the low mass and
long orbital periods of these companions would make the
hot Jupiters appear singular.
(iii) Hot Jupiters with cold Jupiter companions – Similar to
the lonely hot Jupiters, planetary systems that contain both
hot and cold Jupiters tend to arise from solids-rich discs.
Hot Jupiters form early in the disc lifetime and migrate
close to the central star, whilst late forming giant planets
have insufficient time to migrate into the inner system,
retaining long orbital periods as cold Jupiters. Typically,
lower mass planets are found to occupy the space between
the hot and cold Jupiters. Examples of these systems are
shown in Fig. 10 by the prefix 'MJ' (multiple Jupiters),
showing the diversity in planetary compositions in these
systems.
(iv) Cold Jupiters with low-mass companions – When there
is sufficient solid material in the disc, we find that giant
planets can form simultaneously with inner systems of low
mass planets. The late formation of giant planets enables
them to remain as cold Jupiters at the end of the disc
lifetime, whilst interior low-mass planets slowly accrete
and migrate into the inner disc regions, becoming an inner
system of
low-mass planets, occasionally in a compact
configuration. This architecture is similar to that found
in the Solar System, and we note that recent analysis of
Kepler light-curves indicates the existence of long period
giant planets orbiting stars with known compact low mass
systems, similar to our simulated cold Jupiters with short
period low mass companions (Uehara et al. 2016; Kipping
et al. 2016). This planetary system architecture is shown by
systems with the prefix 'CJ' (cold Jupiters) in Fig. 10.
4 DISCUSSION AND CONCLUSIONS
We have presented the results of N-body simulations cou-
pled with prescriptions for planetary migration, accretion of
gaseous envelopes, self-consistent evolution of a viscous disc
with an inner magnetospheric cavity and disc removal by a
photoevaporative wind on multi-Myr time scales. A new ad-
dition, not considered in our previous simulations (Coleman
& Nelson 2014, 2016), is radial structuring of the disc due
to variations in the viscous stresses, leading to the forma-
tion of persistent planet traps at large orbital radii from the
star. The main results from our study can be summarised
as follows:
(i) Radial structuring of the disc allows gas giant planets
to form. Protoplanets and planetesimals become trapped
at the outer edges of the radial structures, due to strong
corotation torques and positive pressure gradients, respec-
tively. Giant planet cores capable of accreting gaseous en-
velopes are able to form due to efficient accretion of plan-
etesimals/boulders by planetary embryos. Out of 288 simu-
lations, 132 surviving giant planets were formed by having
their cores trapped by radial structures. The final periods de-
pend on the time and location of formation, as discussed in
Coleman & Nelson (2014), where early forming giant planets
became hot Jupiters, and late forming giant planets remain
as cold Jupiters.
(ii) When analysing the effects of changing specific parame-
ters, we identify the following trends:
– In solid-poor simulations (low disc mass and metallicity)
no giant planets are formed, as there is insufficient solid ma-
terial to form giant planet cores. This is in agreement with
the observations of Fischer & Valenti (2005) and Santos et al.
(2004), where giant planets are preferentially found around
metal-rich stars.
– When the planetesimal radii are large (≥ 1km), giant plan-
ets are unable to form except in the most solids-rich envi-
ronments. Giant planet formation is strongly favoured in
models where the primary feedstock of planetary growth is
in the form of small 100 m sized planetesimals or 10 m sized
boulders. 95% of the giant planets that formed did so in
simulations with small boulders/planetesimals. None were
formed in models with 10 km planetesimals.
– We find that discs with higher viscosity form fewer giant
planets than low viscosity discs, and the ratio of hot to cold
Jupiters in higher viscosity discs is smaller than in lower
viscosity discs. These effects are entirely due to the shorter
disc lifetimes associated with higher viscosity.
(iii) Multiple giant planets are able to form when there is
sufficient solid material. This occurred in numerous simu-
lations with high disc masses and metallicities, resulting in
systems with multiple cold Jupiters, or a hot Jupiter with
cold Jupiter companions. The survival rate of warm Jupiters
(those with periods between 10–100 d) also depends on the
presence of outer giant companions. Outer giant planet com-
panions can stem the flow of gas into the inner system, re-
ducing the migration rate of planets in the inner system and
allowing them to survive at longer periods than if there were
no exterior giant planets.
(iv) Our simulations reproduce the giant planet period valley
between 10 and 100 days that is seen in the observed period
distribution of giant planets. Our analysis shows that this
arises because of the inclusion of disc removal by photoevap-
oration in our simulations. The launching of a photoevapo-
rative wind causes the disc to empty from the inside out at
the end of its lifetime, causing the migration of planets to
stall at periods > 100 days, an effect that has been discussed
previously by Alexander & Pascucci (2012) and Ercolano &
Rosotti (2015).
(v) Our simulations do not reproduce the broad eccentricity
distribution of the observed giant exoplanets, and this is ap-
parently because our multiple giant planet systems are too
well separated to undergo dynamical instabilities that lead
to the formation of eccentric orbits. We note, however, that
our use of a simple model for damping the eccentricity of gap
forming planets in the presence of the gas disc may also bias
the simulations towards producing low eccentricity systems.
A definitive conclusion about the ability of our models to
form a population of eccentric giants can only be made once
an improved prescription for this has been implemented. A
further point that is worth making is that systems of mul-
tiple giant planets form in the simulations when the system
metallicity is high (as described above). Assuming that the
primary mechanism leading to the observed giant exoplan-
ets attaining their eccentric orbits was dynamical instability
in multiplanet systems (possibly on time scales much longer
than the formation time scales that we have considered), we
note that this (not unexpected) correlation between metal-
licity and the multiplicity of giant planets that form in the
simulations may also explain the positive correlation that
exists between eccentricity and stellar metallicity for giant
MNRAS 000, 000–000 (0000)
exoplanets discovered by radial velocity surveys. 1 We note
that this correlation has also been pointed out by Dawson
& Murray-Clay (2013).
(vi) Numerous compact systems of super-Earths and Nep-
tunes were formed in the simulations, with formation histo-
ries similar to those discussed in Coleman & Nelson (2016). If
there was sufficient solid material, long period giant planets
also formed in the same simulations as the compact systems
of super-Earths/Neptunes.
The simulations we have presented here show that giant
planets can form in discs containing radial structures that
act as planet traps, while the combination of magnetospheric
cavities and photoevaporative winds creates two populations
of giant planets: hot Jupiters and cold Jupiters. It is likely
that in more realistic discs, the location, size and evolution
of radial structures will be quite different from what we have
examined in this work. Running a full parameter study on
the effects of radial structures in protoplanetary discs, how-
ever, goes beyond the scope of this study, which is intended
to be a proof of concept rather than an exhaustive survey of
parameter space.
Whilst this and recent work (Coleman & Nelson 2016)
can more or less explain some of the diversity observed in
exoplanetary system architectures, it is by no means com-
plete. Further improvements to the model are required to
enhance the accuracy and realism currently provided by sim-
ple assumptions. In future work we will aim to include the
following improvements:
(i) A full collision and fragmentation model, so that the
outcomes of planet-planet and planet-planetesimal collisions
can be accurately modelled, instead of the current assump-
tion of perfect mergers. In addition to accounting for effects
arising from impacts between solid bodies, it will also be
important to incorporate a model for the post-collision evo-
lution of gaseous envelopes (Liu et al. 2015a,b).
(ii) Incorporating a more realistic migration model that
takes into account 3D effects (Fung et al. 2015), the in-
fluence of planet luminosity (Ben´ıtez-Llambay et al. 2015)
and dynamical torques arising from the planet's migration
(Paardekooper 2014; Pierens 2015).
(iii) Calculation of gas envelope accretion using self-
consistent calculations that include the effects of changing
local disc conditions and planetesimal accretion rates, rather
than using fits to the Movshovitz et al. (2010) models that
strictly apply only to planets at fixed locations in a non-
evolving disc with specifically prescribed planetesimal ac-
cretion rates.
(iv) Incorporating fits to MHD simulations so that disc ra-
dial structures arising from zonal flows and transitions be-
tween magnetically active and dead zones can be included
in a more realistic fashion.
(v) Incorporating an improved model for the eccentricity
evolution of gap forming planets. At present the model sim-
ply assumes eccentricity damping on a time scale of ∼ 10
orbital periods, independent of the mass contained in the
disc (although the damping is removed after disc removal).
This biases our systems towards low eccentricity, and pre-
1 This correlation may be seen by plotting eccentricity versus
stellar metallicity using the data on radial velocity planets avail-
able at exoplanets.org
MNRAS 000, 000–000 (0000)
Giant planet formation in structured discs
17
vents us from properly assessing whether or not the sim-
ulations that we have presented are able to reproduce the
eccentricity distribution of the observed giant exoplanets.
Finally, it is important to emphasise that this study is
not an exercise in population synthesis, as we have not cho-
sen initial conditions (e.g. disc masses, disc lifetimes) from
observationally motivated distribution functions. Instead we
have simply sought to examine what conditions are required
to form giant planets whose gross orbital characteristics are
similar to the observed distribution. Once our model be-
comes more sophisticated we will examine whether or not it
is capable of reproducing the observed distribution of exo-
planets with appropriate choices of initial conditions.
REFERENCES
ALMA Partnership et al., 2015, ApJ Letters, 808, L3
Alexander R. D., Armitage P. J., 2009, ApJ, 704, 989
Alexander R. D., Pascucci I., 2012, MNRAS, 422, 82
Alibert Y., et al., 2006, A&A, 455, L25
Andrews S. M., et al., 2016, ApJ Letters, 820, L40
Bai X.-N., Stone J. M., 2014, ApJ, 796, 31
Balbus S. A., Hawley J. F., 1991, ApJ, 376, 214
Becker J. C., Vanderburg A., Adams F. C., Rappaport S. A.,
Schwengeler H. M., 2015, ApJ Letters, 812, L18
Ben´ıtez-Llambay P., Masset F., Koenigsberger G., Szul´agyi J.,
2015, Nature, 520, 63
B´ethune W., Lesur G., Ferreira
J.,
2016, preprint,
(arXiv:1603.02475)
Bitsch B., Kley W., 2011, A&A, 536, A77
Bitsch B., Johansen A., Lambrechts M., Morbidelli A., 2015a,
A&A, 575, A28
Bitsch B., Lambrechts M., Johansen A., 2015b, A&A, 582, A112
Cassan A., et al., 2012, Nature, 481, 167
Chambers J. E., 1999, MNRAS, 304, 793
Coleman G. A. L., Nelson R. P., 2014, MNRAS, 445, 479
Coleman G. A. L., Nelson R. P., 2016, MNRAS, 457, 2480
Cossou C., Raymond S. N., Pierens A., 2013, A&A, 553, L2
Cresswell P., Nelson R. P., 2008, A&A, 482, 677
Crida A., Morbidelli A., Masset F., 2006, Icarus, 181, 587
Cumming A., Butler R. P., Marcy G. W., Vogt S. S., Wright J. T.,
Fischer D. A., 2008, PASP, 120, 531
Dawson R. I., Murray-Clay R. A., 2013, ApJ Letters, 767, L24
Desch S. J., Turner N. J., 2015, ApJ, 811, 156
Dipierro G., Price D., Laibe G., Hirsh K., Cerioli A., Lodato G.,
2015, MNRAS, 453, 73
Dittrich K., Klahr H., Johansen A., 2013, ApJ, 763, 117
Duffell P. C., Haiman Z., MacFadyen A. I., D'Orazio D. J., Farris
B. D., 2014, ApJ Letters, 792, L10
Dullemond C. P., Hollenbach D., Kamp I., D'Alessio P., 2007,
Protostars and Planets V, pp 555–572
Durmann C., Kley W., 2015, A&A, 574, A52
Ercolano B., Rosotti G., 2015, MNRAS, 450, 3008
Fendyke S. M., Nelson R. P., 2014, MNRAS, 437, 96
Fischer D. A., Valenti J., 2005, ApJ, 622, 1102
Flock M., Ruge J. P., Dzyurkevich N., Henning T., Klahr H., Wolf
S., 2015, A&A, 574, 68
Fressin F., et al., 2013, ApJ, 766, 81
Fromang S., Nelson R. P., 2006, A&A, 457, 343
Fung J., Artymowicz P., Wu Y., 2015, ApJ, 811, 101
Han E., Wang S. X., Wright J. T., Feng Y. K., Zhao M., Fakhouri
O., Brown J. I., Hancock C., 2014, PASP, 126, 827
Hasegawa Y., Pudritz R. E., 2011, MNRAS, 417, 1236
Hayashi C., 1981, Progress of Theoretical Physics Supplement,
70, 35
Hellary P., Nelson R. P., 2012, MNRAS, 419, 2737
18 G. A. L. Coleman and R. P. Nelson
Huber D., et al., 2014, ApJS, 211, 2
Ida S., Lin D. N. C., 2004, ApJ, 616, 567
Inaba S., Ikoma M., 2003, A&A, 410, 711
Johansen A., Youdin A., Mac Low M.-M., 2009, ApJ Letters, 704,
L75
Kasting J. F., Whitmire D. P., Reynolds R. T., 1993, Icarus, 101,
108
Kipping D. M., et al., 2016, preprint, (arXiv:1603.00042)
Levison H. F., Kretke K. A., Duncan M. J., 2015, Nature, 524,
322
Lin D. N. C., Papaloizou J., 1986, ApJ, 309, 846
Lissauer J. J., et al., 2011, Nature, 470, 53
Liu S.-F., Agnor C. B., Lin D. N. C., Li S.-L., 2015a, MNRAS,
446, 1685
Liu S.-F., Hori Y., Lin D. N. C., Asphaug E., 2015b, ApJ, 812,
164
Lodders K., 2003, ApJ, 591, 1220
Lynden-Bell D., Pringle J. E., 1974, MNRAS, 168, 603
Marzari F., Weidenschilling S. J., 2002, Icarus, 156, 570
Masset F. S., Morbidelli A., Crida A., Ferreira J., 2006, ApJ, 642,
478
Mayor M., Queloz D., 1995, Nature, 378, 355
Mordasini C., Alibert Y., Benz W., 2009, A&A, 501, 1139
Movshovitz N., Bodenheimer P., Podolak M., Lissauer J. J., 2010,
Icarus, 209, 616
Muirhead P. S., et al., 2012, ApJ, 747, 144
Nayakshin S., 2015, MNRAS, 454, 64
Neveu-VanMalle M., et al., 2016, A&A, 586, A93
Okuzumi S., Momose M., Sirono S.-i., Kobayashi H., Tanaka H.,
2015, preprint, (arXiv:1510.03556)
Paardekooper S.-J., 2014, MNRAS, 444, 2031
Paardekooper S.-J., Baruteau C., Crida A., Kley W., 2010, MN-
RAS, 401, 1950
Paardekooper S.-J., Baruteau C., Kley W., 2011, MNRAS, 410,
293
Papaloizou J. C. B., Nelson R. P., 2003, MNRAS, 339, 983
Picogna G., Kley W., 2015, preprint
Pierens A., 2015, MNRAS, 454, 2003
Pinilla P., Birnstiel T., Ricci L., Dullemond C. P., Uribe A. L.,
Testi L., Natta A., 2012, A&A, 538, A114
Rasio F. A., Ford E. B., 1996, Science, 274, 954
Santerne A., et al., 2015, preprint, (arXiv:1511.00643)
Santos N. C., Israelian G., Mayor M., 2004, A&A, 415, 1153
Shakura N. I., Sunyaev R. A., 1973, A&A, 24, 337
Simon J. B., Beckwith K., Armitage P. J., 2012, MNRAS, 422,
2685
Sousa S. G., et al., 2008, A&A, 487, 373
Steinacker A., Papaloizou J. C. B., 2002, ApJ, 571, 413
Udry S., Mayor M., Santos N. C., 2003, A&A, 407, 369
Uehara S., Kawahara H., Masuda K., Yamada S., Aizawa M.,
2016, preprint, (arXiv:1602.07848)
Umebayashi T., Nakano T., 1988, Progress of Theoretical Physics
Supplement, 96, 151
Weidenschilling S. J., 1977, MNRAS, 180, 57
Zhang K., Blake G. A., Bergin E. A., 2015, ApJ Letters, 806, 7
Zhang K., Bergin E. A., Blake G. A., Cleeves L. I., Hogerheijde
M., Salinas V., Schwarz K. R., 2016, ApJ Letters, 818, L16
Zhu Z., Stone J. M., Rafikov R. R., Bai X.-n., 2014, ApJ, 785,
122
MNRAS 000, 000–000 (0000)
|
1210.4827 | 2 | 1210 | 2013-01-31T19:25:33 | A Possible Divot in the Size Distribution of the Kuiper Belt's Scattering Objects | [
"astro-ph.EP"
] | Via joint analysis of a calibrated telescopic survey, which found scattering Kuiper Belt objects, and models of their expected orbital distribution, we measure the form of the scattering object's size distribution. Ruling out a single power-law at greater than 99% confidence, we constrain the form of the size distribution and find that, surprisingly, our analysis favours a very sudden decrease (a divot) in the number distribution as diameters decrease below 100 km, with the number of smaller objects then rising again as expected via collisional equilibrium. Extrapolating at this collisional equilibrium slope produced enough kilometer-scale scattering objects to supply the nearby Jupiter-Family comets. Our interpretation is that this divot feature is a preserved relic of the size distribution made by planetesimal formation, now "frozen in" to portions of the Kuiper Belt sharing a "hot" orbital inclination distribution, explaining several puzzles in Kuiper Belt science. Additionally, we show that to match today's scattering-object inclination distribution, the supply source that was scattered outward must have already been vertically heated to of order 10 degrees. | astro-ph.EP | astro-ph | A Possible Divot in the Size Distribution of the Kuiper Belt's
Scattering Objects
C. Shankman, B. J. Gladman,
Department of Physics and Astronomy, University of British Columbia, 6224 Agricultural
Road, Vancouver, BC, V6T 1Z1
N. Kaib
Department of Physics and Astronomy, Queens University, Canada
Herzberg Institute of Astrophysics, National Research Council of Canada, Victoria, BC,
J.J. Kavelaars
Canada
and
J.M. Petit
Institut UTINAM, CNRS-Universit´e de Franche-Comt´e, Besan¸con, France
Received
;
accepted
3
1
0
2
n
a
J
1
3
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
7
2
8
4
.
0
1
2
1
:
v
i
X
r
a
-- 2 --
ABSTRACT
Via joint analysis of a calibrated telescopic survey, which found scattering
Kuiper Belt objects, and models of their expected orbital distribution, we explore
the scattering-object size distribution. Although for D >100 km the number of
objects quickly rise as diameters decrease, we find a relative lack of smaller ob-
jects, ruling out a single power-law at greater than 99% confidence. After study-
ing traditional "knees" in the size distribution, we explore other formulations and
find that, surprisingly, our analysis is consistent with a very sudden decrease (a
divot) in the number distribution as diameters decrease below 100 km, which then
rises again as a power-law. Motivated by other dynamically hot populations and
the Centaurs, we argue for a divot size distribution where the number of smaller
objects rises again as expected via collisional equilibrium. Extrapolation yields
enough kilometer-scale scattering objects to supply the nearby Jupiter-Family
comets. Our interpretation is that this divot feature is a preserved relic of the
size distribution made by planetesimal formation, now "frozen in" to portions of
the Kuiper Belt sharing a "hot" orbital inclination distribution, explaining sev-
eral puzzles in Kuiper Belt science. Additionally, we show that to match today's
scattering-object inclination distribution, the supply source that was scattered
outward must have already been vertically heated to of order 10◦.
Subject headings: comets: general -- Kuiper belt: general
-- 3 --
1.
Introduction
Measurements of the Kuiper Belt's size distribution (number at each diameter D)
constrain accretional processes at planet formation and, potentially, subsequent collisional
or physical evolution. Because astronomers observe brightnesses rather than D, object
absolute magnitudes H are tabulated as the observable proxy for the size distribution.
Collisional and accretional theories suggest exponential forms for the N(H) distribution.
A differential number distribution of the form dN/dH ∝ 10αHg with a logarithmic 'slope'
α corresponds to a power-law D distribution dN/dD ∝ D−(5α+1). Although power-law
D distributions provide acceptable fits to Kuiper Belt surveys over spans of a few
magnitudes in H, departures from single power-laws are necessary over larger H ranges
(Jewitt et al. 1988; Gladman et al. 2001; Bernstein et al. 2004; Fuentes & Holman 2008;
Fraser & Kavelaars 2008). For the steep (α=0.8 -- 1.2) distributions seen in the Kuiper Belt,
detections are dominated by objects near the largest H magnitude (smallest size) visible
in a given survey. An α > 0.6 slope cannot continue as H → ∞ (D → 0 km) or the total
mass diverges; thus a slope change (generically called a break) is required. Evidence of
such a break now exists for trans-Neptunian objects (TNOs) in the main Kuiper Belt (at
distances d ≃ 38 -- 46 AU), both near the sensitivity limit of ground-based telescopic surveys
(Fraser & Kavelaars 2008; Fuentes & Holman 2008) (reaching Hg ∼9 -- 10 at 40 AU) and
from deeper HST (Bernstein et al. 2004) observations (Hg ∼13 at 40 AU). This break has
been modelled as a gradual transition to a smaller value of α, a "knee".
Probing a break is difficult because small TNOs are faint. This problem is reduced
when observing the scattering objects (SOs); these are mostly TNOs with perihelia q ≤
35 AU (see below) and thus smaller objects are detectable while near the Sun.
At any time some SOs are only d = 20 -- 30 AU away, allowing a 4-m telescope, in
excellent conditions, to detect objects down to Hg ∼ 12. For a monotonically increasing
-- 4 --
number distribution N(Hg), the abundant small objects at the observable volume's
innermost edge should dominate the detected sample. This is not what our survey found
(Fig. 1), necessitating a relative lack of small SOs.
We discarded a sudden (ad-hoc) albedo change, as it would produce a gap in H-space,
not a drop, which does not match the observations; the needed change is a sudden lack of
Hg > 9 (and therefore small) SOs.
2. Models
Several models of the SO orbital element distribution were exposed to the calibrated
observational biases of the Canada France Ecliptic Plane Survey (CFEPS) in order to
quantitatively constrain the intrinsic N(Hg) distribution. Drawing SOs from an orbital
distribution model, and selecting Hg from a candidate N(Hg) distribution, the CFEPS
survey simulator (Jones et al. 2006) determines each object's observability and produces a
set of "simulated detections" expected from the model.
Two different SO orbital models are from a modified version of Kaib et al. (2011b)
(henceforth KRQ11). KRQ11 focuses on the effects of solar migration in the Milky Way
on Oort Cloud structure. While this is not the focus of the current work, we can use the
KRQ11 control calculations, which assume an unchanging local galactic environment.
To test the sensitivity of our results to the dynamical context, we performed the same
analysis on an independent model. Gladman & Chan (2006) modelled the scattering of
objects in an initial Solar System having an additional planet of order Earth mass. As
previously reported (Petit et al. 2011) this model also (perhaps surprisingly) satisfactorily
represents the current SO (a, q) distribution, although too "cold" in inclinations. In fact,
this model and the cold KRQ11 model produced very similar results, showing that our
-- 5 --
conclusions are mostly insensitive to the assumed Solar System history. Objects currently
in the Centaur and detectable SO region (mostly a <200 AU) have, unsurprisingly, almost
forgotten their initial state except for the the inclination distribution; the current SO orbital
distribution is not diagnostic of the number and position of the planets early in the Solar
System's history.
3. Observations
CFEPS provided a set of detections of outer Solar System objects in a precisely
calibrated survey (Jones et al. 2006; Kavelaars et al. 2009) whose pointing history, detection
efficiency, and tracking performance were recorded. The final set of TNO detections (with
full high-precision orbits) and the fully calibrated pointing history make up the L7 release
(Petit et al. 2011). This absolute calibration of CFEPS allows a model of the present orbital
(and size) distribution of to be passed through the CFEPS Survey Simulator, yielding a
set of simulated detections whose orbital and Hg distributions can be compared to the real
detections.
The three models provide orbital distributions of all TNOs. The "scattering" TNOs
are then selected out of the final 10 Myr stage of the model integrations using the
criteria: variation of a >1.5 AU in semimajor axis during 10 Myr, with a < 1000 AU
(Morbidelli et al. 2004; Gladman et al. 2008). Historically, a simple q cut was used to
isolate the "scattered disk" (Duncan & Levison 1997; Luu et al. 1997; Trujillo et al. 2000),
which has serious disadvantages when trying to discuss the cosmogony, as there is a nearly
impossible distinction between implanted (and thus scattered) and the original Kuiper Belt
population (if any). Perihelion divisions also undesirably includes resonant objects and
most inner main-belt TNOs.
-- 6 --
The CFEPS SO sample consists of 9 objects (Table 1), supplemented by two SOs
discovered in a high-latitude extension survey (covering ≃ 470 sqdg in 2007 -- 2008, extending
up to 65◦ ecliptic latitude), which was fully calibrated in the same way as CFEPS.
To characterize the form of the N(Hg), we introduce a novel formulation, allowing for
the exploration of distributions with knees and divots (a sudden drop in the differential
number of objects followed by a recovery). We parameterised the Hg distribution (Fig. 2A)
with the fixed slope αb = 0.8 (see below) for SOs brighter than a break at Hg = 9
(D ≃100 km), allowed an adjustable slope αf for fainter objects, and an adjustable contrast
c ≥ 1.
The Hg-magnitudes are drawn from one of three types of distributions:
(1) a single exponential of logarithmic slope α, (2) N(Hg) with a knee (contrast c =1). That
is, one slope αb for SOs with Hg < Hknee and αf for Hg > Hknee, where N(Hg) is continuous
across the knee at Hknee and negative slopes αf are allowed as suggested (Bernstein et al.
2004), and (3) one slope αb to a divot at Hdivot, which is a sudden drop in differential
number by a factor c, with a potentially different slope αf beyond the cliff at H = Hdivot.
Although in reality the discontinuity is unlikely to be an instantaneous drop, our data do
not merit trying to constrain the values of the expected steep negative slope and small
extent over which it drops; collisional models (Fraser 2009; Campo Bagatin & Benavidez
2012) do show collisional divots where the drop occurs over D ranges of factors < 2 (a few
tenths of magnitude in Hg).
In principle there are four parameters: αb, αf , Hdivot, and c (Fig. 2). For Hg < 9 a
single power-law of αb ≃ 0.8 does indeed match our detections; we elected to fix this slope
at that value with the unifying philosophy that all the hot transneptunian populations
share this same hot slope; αb = 0.8 matches both the hot Classical belt measured down to
Hg ≃ 8.0 (Petit et al. 2011; Fraser & Kavelaars 2008; Fuentes & Holman 2008), and to the
-- 7 --
3:2 resonators measured down to Hg ≃ 9.0 (Gladman et al. 2012). Our detections require a
transition around Hg =9 -- 10 to explain the relative lack of small detections; we thus fixed
the knee/divot for our analysis at Hg = 9 (slightly larger than D=100 km for 5% g-band
albedo). This leaves only two free parameters: the contrast c at the divot and the slope αf
for absolute magnitudes fainter than the divot/knee.
To assess a match, the Anderson-Darling (AD) statistic is calculated between our
11-object sample and the distribution of simulated detections from the model, for each
orbital parameter. An AD significance level of < 5% rejects the hypothesis that the real
SO observations could be drawn from the simulated detections at the 95% confidence level
(for that orbital parameter). To retain a model, we required that none of the q, d, i, and
Hg distributions are rejectable at > 95% confidence.
4. Absolute Magnitude Distribution
The observational bias is strong (Fig. 1) , but when accurately calibrated allows us to
constrain the Hg distribution's form. Single power-laws predict significantly more close-in
detections than were seen by CFEPS; for a slope of α = 0.8, roughly half of the expected
detections (Fig. 1 D's blue dashed curve) should have a distance at detection d < 23 AU,
which is the closest real SO in our sample. The observationally biased models predict that
the majority of detected SOs would have orbits with q <20 AU at d = 20 -- 25 AU and be
small (Hg > 9 or D ≤ 100 km) objects, in contrast to our detections, which demonstrates
that our observations are sensitive beyond the break. When confined to q > 25AU (where
objects must be large to be seen) the orbital models provide good matches, however
extensions to smaller distance fail when using a single power-law, pointing to a breakdown
arising from the the assumed N(Hg).
-- 8 --
We rule out a single power-law of slope 0.8 at 99% confidence, and can rule out all
single power-laws with slope between 0 and 1.2 at 95% confidence. Slopes of 0.5 and 0.6
are not rejectable across the whole distribution, but are rejectable (95% confidence) when
the distribution is considered in both Hg > 9 and Hg < 9 subsets; we demand these work
because the steep slopes measured for other hot populations match our Hg < 9 detections
well, and a shallower slope is erroneously found by measuring across a divot feature when
requiring a single slope (see below).
Our relatively small sample is powerful because our detected SOs span the break and,
when coupled with the precise CFEPS calibration, allows the non-detection of Hg =10-12
SOs (several magnitudes past the divot) to provide a strong constraint on N(Hg). Down
to this limit, CFEPS detected moving objects as close as 20 AU with no rate of motion
dependence. Because our orbits are accurate, we can separate the SOs from the other hot
populations, and use a dynamical model specific to the SOs.
All of our N(Hg) cases have the obvious and previously-known problem that the
model's orbital inclinations are mostly lower than the true population's (Petit et al. 2011;
Gladman et al. 2012), even for the cases where the N(Hg) otherwise provides a good match.
For example, Fig. 1 shows a divot (c ≃ 6, αf = 0.5) producing a good match between
the model's expected detections (green curve) and the real SO sample (red), excepting
the i problem. Models (Levison et al. 2008) which scatter out a cold TNO population
(from d < 30 AU with initial inclination distribution widths σi ≤ 6◦) to eventually
form today's hot population produce current TNO populations where too many low-i
detections are expected in observational surveys (Fig. 1B). This is part of growing evidence
that the original planetesimal disk supplying today's high-i objects must have already
been vertically excited before being scattered out (Petit et al. 2011; Gladman et al. 2012;
Brasser & Morbidelli 2012), which has strong cosmogonic implications for an extended
-- 9 --
quiescent phase of the early Solar System (Levison et al. 2008). We therefore computed a
new SO model with a hotter (σi ≃ 12◦) initial disk; this provides an excellent match with
today's SO i-distribution (see Fig. 3) and we constrain N(Hg) below using this model.
To constrain the size distribution, a grid of possible divot contrasts and post-divot
slopes was explored. Fig. 4 shows acceptability levels for the range of explored parameter
pairs (c, αf ). A single power-law of α = 0.8 (blue star Fig. 4) has < 1% probability. We
are left with a range of acceptable parameter space, including knee (c=1) and divot (c >1)
scenarios; we further constrain N(Hg) by looking to other Kuiper Belt populations.
The so-called hot Kuiper Belt populations (the hot main belt, inner belt, resonant, and
detached TNOs) share an i distribution half-width of roughly 15◦ (Petit et al. 2011) with
the SOs, suggesting a cosmogonic link. In analyses of deep luminosity functions dominated
by hot main-belt detections, the common conclusion (Bernstein et al. 2004; Fuentes et al.
2010) was that for magnitude g > 25 the slope must break to a faint αf < 0.3 value or even
become negative in order to explain the lack of detections in the following few magnitudes;
beyond this no data exists for the main Kuiper Belt. For SOs and their companion objects
(Centaurs), however, many Hg ≫ 9 objects are known from wide-field surveys, mostly
detected at d < 20 AU. In fact, measurements of Jupiter Family Comets (JFCs) in the
Hg ≈14 -- 17 range give slopes αf ≃ 0.5 ± 0.1 (see Table 6 of Solontoi et al. (2012)). These
two arguments mean the SO distribution cannot remain at αf < 0.3. leading us to discard
knees to negative slopes. A divot can explain both a relative lack of objects beyond the
break and the eventual recovery necessary to provide the JFCs. A divot also motivates the
negative slopes measured, as a realistic divot will take the form of a decrease at the break,
rather than the sharp discontinuity we use. We prefer the divot solution with αf =0.5
and c ≃6 (green star Fig. 4) which matches the observations and allows for a single slope
αf = 0.5 from the divot out to the Hg >14 Jupiter Family comets whose slope is near the
-- 10 --
collisional equilibrium value (O'Brien & Greenberg 2005).
5. External Arguments
As the hot populations have similar colours and D > 100-km size distributions, it seems
likely that they were all transplanted to join a pre-existing cold Kuiper Belt (Petit et al.
2011) during a common event early in the Solar System's history, and would thus logically
share the same divot and small D distribution. Such a transplant process can successfully
implant TNOs in the stable Kuiper Belt (Levison et al. 2008; Batygin et al. 2011), although
the resonant population ratios and i distribution are problematic (Gladman et al. 2012).
We thus look for evidence of such a feature in other hot populations.
5.1. The Neptune Trojans
A search for Neptune Trojans (Sheppard & Trujillo 2010) provided significant evidence
that an α ∼ 0.8 power-law cannot continue for D < 100-km Trojans; a divot was
not apparent because Trojans significantly smaller were not detected. The dispersed
Trojan inclination distribution (Sheppard & Trujillo 2006), although not yet precisely
measured, links these objects to all the other resonant populations (Gladman et al. 2012).
Sheppard & Trujillo (2010) used Neptune Trojan searches to argue that beyond mR ≃23
(corresponding to Hg ≃ 9) there was an absence of Trojans due to non-detections, and
thus smaller Trojans were missing. Assuming that the Trojans and other resonant TNOs
were implanted from a scattering population, and thus share the same size distribution, we
confirmed that our divot N(Hg) matches the lack of D <100 km Neptune Torjan detections
in the Sheppard & Trujillo (2010) surveys.
Given our analysis, the conclusion would not be that small Neptune Trojans are
-- 11 --
"missing", but rather that the sudden drop results in the population fainter than the
divot not recovering in on-sky surface density until at least Hg > 11, by which point the
deepest survey lacked the sensitivity to detect them. If correct, detection of several small
(Hg > 11) Trojans requires surveying ∼ 100 square degrees to 26th magnitude at the
correct elongation.
5.2. Hot Populations
A recent deep telescopic survey (Fraser et al. 2010) estimated α ≃ 0.40 ± 0.15 (within
error of our preferred αf = 0.5) from the apparent-magnitude distribution for "close"
(30 < d < 38) TNOs (orbits were not obtained). These distances are dominated by several
hot populations, but the measurement is shallower than the usual hot population slope of
0.8. We calculated Hg magnitudes for the Fraser et al. (2010) detections and find that due
to the survey's depth, this sample is dominated by Hg > 9 TNOs and thus would measure
the post-divot slope.
6. Feasibility of a Divot
A primordial size-distribution wave at small sizes (D = 2 km) could propagate (Fraser
2009) to D ∼100 km in a dynamically hot (∆v = 2 km/s) collisional environment after
500 Myr. Alternately, recent modeling of planetesimal creation (Johansen et al. 2007;
Morbidelli et al. 2009), suggests that the protosolar nebula may only have produced
planetesimals larger than a certain critical diameter, in which case the αb = 0.8 slope and
the D ∼ 100 km divot size are set by planetesimal formation physics; smaller objects
appear only later due to collisional fragmentation. These scenarios match our results,
where one interprets the hot population's N(Hg) to have been "frozen" when suddenly
-- 12 --
transplanted (scattered) from a denser region nearer the Sun to the large volume it now
occupies, ending the collisional evolution. An exciting prospect is that the divot directly
records a preferential D that planet building produced in the solar-nebula region where
the hot TNOs originally formed, and that the divot's depth (which could easily range from
c=2 -- 30) measures the integrated collisional evolution (depending on both the duration of
the pre-scattering phase and the random speeds present). An initial distribution with no
D <100 km TNOs was shown (Campo Bagatin & Benavidez 2012) to evolve into a divot
with c ∼ 20 and αf ≃ 0.5, in the dynamical environment of the Nice model; such a 500-Myr
quiescent phase (Gomes et al. 2005) allows a divot to form but the evidence we find for
a higher-i early phase may argue instead for a much shorter and more intense collisional
environment.
Our divoted N(Hg) produces a cumulative distribution (Fig. 2 B) with a shallow
plateau for Hg=9 -- 12, similar to that deduced for the hot population and SOs in deep HST
observations (Bernstein et al. 2004; Volk & Malhotra 2008) and estimated for scattering
impactors of the saturnian moons (Minton et al. 2012). Our estimate of 2 × 106 SOs with
Hg < 13 and a slope of αf = 0.5 extrapolates to ∼ 2 × 109 SOs with Hg <18, providing
a sufficient number (Duncan & Levison 1997; Volk & Malhotra 2008) of SOs to feed the
Jupiter Family Comets, while satisfying the observed plateau.
Single power-laws that fit the Hg < 9 SOs fail when extended to smaller objects. Our
novel divot parameterisation (Fig. 2) matches our data and would simultaneously explain
the puzzles of the JFC source, the "missing" Neptune Trojans, and the known rollover in
the Kuiper Belt's luminosity function. To better constrain the form of the break, a new
survey must find and determine orbits for ∼10 SOs from 10 -- 30 AU; this requires discovery
(and tracking over several degrees of arc) targets moving up to 15"/hr by observing ∼200
sq. deg. to 24th magnitude.
-- 13 --
Facilities: CFHT, NRC (HIA)
-- 14 --
REFERENCES
Bernstein, G.M. et al. 2004, AJ, 128, 1364
Brasser, R. & Morbidelli, A. 2012 Asteroids, Comets, and Meteors meeting, abstract.
Batygin, K., Brown, M. & Fraser, W. 2011, ApJ, 13, 738
Campo Bagatin & A.,Benavidez, P. 2012, MNRAS, 423, 1254
Duncan, M. & Levison, H.F. 1997, Science, 276, 1670
Elliot, J., Kern, S.D. & Clancy, K.B. 2005, AJ, 129, 1117
Fraser, W. 2009, ApJ, 706, 119
Fraser, W., Brown, M. & Schwamb, M. 2010, Icarus, 210, 944
Fraser, W. & Kavelaars J.J. 2008, Icarus, 198, 452
Fraser, W. & Kavelaars, J.J. 2009, AJ, 137, 72
Fuentes C., & Holman, M. 2008, AJ, 136, 83
Fuentes, C., Holman, M., Trilling, D. & Protopapas, P. 2010, ApJ, 722, 1290
Gladman, B. et al. 2001, AJ, 122, 1051
Gladman, B. & Chan, C. 2006, ApJ, 643, L135
Gladman, B., Kavelaars, J.J., Petit, J.-M., et al. 2009, ApJ, 697, L91
Gladman, B., Lawler, S., Petit, J.-M, et al. 2012, AJ, 144, 23
Gladman, B., Marsden, B.G. & Vanlaerhoven, C. 2008, in The Solar System Beyond
Neptune 43
-- 15 --
Gomes, R., Levison,H.F., Tsiganis & K., Morbidelli, A. 2005, Nature, 435, 466
Jewitt, D., Luu, J., & Trujillo, C. 1988, AJ115, 2125
Johansen, A., Oishi, J., Mac Low, M.-M., et al. 2007, Nature, 448, 1022
Jones, R.L., Gladman, B., Petit, J.-M., et al. 2006, Icarus, 185, 508
Kaib, N., Quinn, T. & Brasser, R. 2011, ApJ, 141, 3
Kaib, N., Roskar, R. & Quinn, T. 2011, Icarus, 215, 491
Kavelaars, J.J., Jones, R.L, Gladman, B., et al. 2009, AJ, 137, 491
Levison, H.F., Dones, L. & Duncan, M. 2001, AJ, 121, 2253
Levison, H., Morbidelli, A., Van Laerhoven, C., Gomes, R. & Tsiganis, K. 2008, Icarus, 196,
258
Luu, J., Marsden, B., Jewitt, D., et al. 1997, Nature, 387, 573
Minton, D., Richardson, J., Thomas, P., Kirchoff, M. & Schwamb, M. 2012, in ACM Conf.
Sess 551
Morbidelli, A., Bottke, W.F., Nesvorn´y, D. & Levison, H. 2009, Icarus, 204, 448
Morbidelli, A., Emel'yanenko, V. & Levison, H.F. 2004, MNRAS, 355, 935
O'Brien, D. & Greenberg, R. 2005, Icarus, 178, 179
Petit, J.-M, Kavelaars, J.J., Gladman, B., et al. 2011, AJ, 142, 131
Sheppard, S. & Trujillo, C. 2006, Science, 313, 511
Sheppard, S. & Trujillo, C. 2010, ApJ, 723, L233
-- 16 --
Solontoi, M., Ivezi´c, Z., Juri´c, M. et al. 2012, Icarus, 218, 571
Volk, K. & Malhotra, R. 2008, ApJ, 687, 714
Trujillo, C., Jewitt, D., & Luu, J. 2000, ApJ, 529, 103
This manuscript was prepared with the AAS LATEX macros v5.2.
-- 17 --
A
20 40 60 80 100 120 140 160
Semimajor axis a(AU)
D
B
E
10
20
30
Inclination i( ◦ )
C
F
40
0.08
0.06
0.04
0.02
q
a
15
20
25
30
Zoom-in a,q(AU)
10 15 20 25 30 35 40
Perhelion q(AU)
20 25 30 35 40 45
Detection distance d(AU)
7
8
9
10
11
Absolute H-magnitude Hg
n
o
i
t
c
a
r
f
e
v
i
t
a
u
m
u
c
l
n
o
i
t
c
a
r
f
e
v
i
t
a
u
m
u
c
l
1.0
0.8
0.6
0.4
0.2
0.0
1.0
0.8
0.6
0.4
0.2
0.0
Fig. 1. -- Cumulative distributions for 5 quantities, comparing the observed objects (red
staircase) with the initially cold model's simulated detections. The black dash-dot curves
show the orbital model's intrinsic a, i, q, and d distributions. When coupled to two different
N(Hg) distributions, the biases produce the differing predictions for the detections. A single
power-law slope of α = 0.8 (blue dashed) is rejectable at > 99% confidence in d, q and Hg.
Contrastingly, our preferred divot N(Hg) (green curve, see Fig. 2) provides vastly better
matches, although both produce too many low-i detections.
-- 18 --
Fig. 2. -- Histograms of the N(Hg) distribution for our preferred divot solution. The vertical
axes show the total SO numbers using the absolute CFEPS calibration (Petit et al. 2011).
A: The differential distribution (solid green), with an extrapolated α = 0.8 beyond Hg = 9
(dashed blue). The contrast c ≃ 6 is the ratio of differential number on either side of the
divot. B: The cumulative version. For Hg >13 the cumulative N(< Hg) has reached slope
αf after the flattened region following the divot.
-- 19 --
AD = 77.1
io ≃12 ◦
AD = 0.28
io ≃1 ◦
10
20
30
40
orbital inclination i (degrees)
1.0
0.8
0.6
0.4
0.2
0.0
0
i
<
n
o
i
t
c
a
r
f
e
v
i
t
a
u
m
u
c
l
Fig. 3. -- The hot (black dashed) and cold (light blue dashed) KRQ11 intrinsic i distributions
produce the biased distributions (solid lines) from our preferred model for comparison with
the CFEPS sample (red). The hot model significantly improves the match because it has
a relative lack (at the current epoch) of low-i SOs to be detected by the survey; the eleven
CFEPS SOs have < 1% probability of being drawn from the initially cold simulation.
-- 20 --
50
35
30
25
15
l
e
v
e
L
y
t
i
l
i
b
a
t
p
e
c
c
A
5
1
10.0
contrast c
100.0
αf
1.2
1.0
0.8
0.6
0.4
0.2
0.0
−0.2
−0.4
1.0
Fig. 4. -- Contour values (colours) computed from a grid (points) in (c, αf ) space, giving
the Anderson-Darling probability of drawing the most rejectable of the d, q, i, and Hg
distributions from the hot KRQ model of today's SOs. White areas indicate when the (c,αf )
pair had <1% probability of coming from the model, and the black contour bounds <5%.
A single power-law (blue, darker star) is rejected. Our favoured model (green, lighter star)
satisfies both (a) αf ≃ 0.5 like known JFCs, (b) α ≤ 0.6 which prevents the extrapolated
mass of small SOs from diverging.
-- 21 --
Designation a (AU) q (AU)
i (deg) d (AU) Hg
L4k09
HL8a1
L4m01
L4p07
L3q01
L7a03
L4v11
L4v04
L4v15
L3h08
HL7j2
30.19
32.38
33.48
39.95
50.99
59.61
60.04
64.10
68.68
159.6
24.60
13.586
26.63
22.33
42.827
44.52
28.73
8.205
31.36
26.31
23.545
29.59
33.41
22.26
6.922
4.575
38.17
46.99
9.5
7.3
8.9
7.7
8.1
7.1
31.64
11.972
26.76
10.0
38.10
13.642
31.85
36.81
14.033
22.95
20.26
15.499
38.45
9.1
9.0
8.0
8.4
133.25
20.67
34.195
37.38
Table 1: CFEPS + extension SO sample
|
1301.2741 | 2 | 1301 | 2013-01-15T07:45:37 | The HARPS search for southern extrasolar planets: XXXIII. New multi-planet systems in the HARPS volume limited sample: a super-Earth and a Neptune in the habitable zone | [
"astro-ph.EP"
] | The vast diversity of planetary systems detected to date is defying our capability of understanding their formation and evolution. Well-defined volume-limited surveys are the best tool at our disposal to tackle the problem, via the acquisition of robust statistics of the orbital elements. We are using the HARPS spectrograph to conduct our survey of ~850 nearby solar-type stars, and in the course of the past nine years we have monitored the radial velocity of HD103774, HD109271, and BD-061339. In this work we present the detection of five planets orbiting these stars, with m*sin(i) between 0.6 and 7 Neptune masses, four of which are in two multiple systems, comprising one super-Earth and one planet within the habitable zone of a late-type dwarf. Although for strategic reasons we chose efficiency over precision in this survey, we have the capability to detect planets down to the Neptune and super-Earth mass range, as well as multiple systems, provided that enough data points are made available. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. finalarx
August 15, 2018
c(cid:13) ESO 2018
3
1
0
2
n
a
J
5
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
4
7
2
.
1
0
3
1
:
v
i
X
r
a
The HARPS search for southern extrasolar planets (cid:63)
XXXIII. New multi-planet systems in the HARPS volume limited sample:
a super-Earth and a Neptune in the habitable zone. (cid:63)(cid:63)
G. Lo Curto1, M. Mayor2, W. Benz3, F. Bouchy2,5, G. H´ebrard5,6, C. Lovis2, C. Moutou4, D. Naef2, F. Pepe2,
D. Queloz2, N.C. Santos7,8, D. Segransan2, and S. Udry2
1 ESO - European Southern Observatory, Karl-Schwarzschild-Strasse 2, 85748 Garching bei Munchen, Germany
2 Observatoire de Gen`eve, Universit´e de Gen`eve, 51 ch. des Maillettes, 1290 Sauverny, Switzerland
3 Physikalisches Institut Universitat Bern, Sidlerstrasse 5, 3012 Bern, Switzerland
4 Laboratoire d'Astrophysique de Marseille, Traverse du Siphon, 13376 Marseille 12, France
5 Institut d'Astrophysique de Paris, UMR7095 CNRS, Universit´e Pierre & Marie Curie, 98bis boulevard Arago, 75014 Paris, France
6 Observatoire de Haute-Provence, CNRS/OAMP, 04870 Saint-Michel-l'Observatoire, France
7 Centro de Astrofisica, Universidade do Porto, Rua das Estrelas, 4150-762 Porto, Portugal
8 Departamento de F´ısica e Astronomia, Faculdade de Ciencias, Universidade do Porto, Portugal
ABSTRACT
The vast diversity of planetary systems detected to date is defying our capability of understanding their formation and evolution.
Well-defined volume-limited surveys are the best tool at our disposal to tackle the problem, via the acquisition of robust statistics of
the orbital elements. We are using the HARPS spectrograph to conduct our survey of ≈ 850 nearby solar-type stars, and in the course
of the past nine years we have monitored the radial velocity of HD103774, HD109271, and BD-061339. In this work we present
the detection of five planets orbiting these stars, with m sin(i) between 0.6 and 7 Neptune masses, four of which are in two multiple
systems, comprising one super-Earth and one planet within the habitable zone of a late-type dwarf. Although for strategic reasons
we chose efficiency over precision in this survey, we have the capability to detect planets down to the Neptune and super-Earth mass
range, as well as multiple systems, provided that enough data points are made available.
Key words. extra-solar planets -- multiple planetary systems -- radial velocity -- HD103774 -- HD109271 -- BD-061339
1. Introduction
After more than 15 years of exo-planets detections (Mayor 1995,
Schneider 2012), we have witnessed an extreme diversity of
planetary systems, but we have not yet detected a system that
could be considered a twin to our own solar system, not even
limited to the giant (and easier to detect) planets. We are left
with the question of whether our system, and with it our planet,
are the norm, or rather the exception in planet formation and evo-
lution. To date, only about one third of the detected planets are
found in confirmed multi-planet systems (Schneider 2012). This
fraction is steadily rising with the increasing amount of available
high-quality data, indicating that we are fighting against a detec-
tion bias. Multi-planet systems are more challenging to detect
because they require many data points to permit the identifica-
tion of the individual signals and because, as is revealed from
the data collected so far, often they involve low-mass planets
in the super-Earth to Neptune mass range (e.g. Lo Curto 2010,
H´ebrard 2010, Lovis 2011, Fischer 2012). The combination of
the low amplitude of these signals with the more complex mo-
Send offprint requests to: G. Lo Curto
e-mail: [email protected]
(cid:63) Based on observations made with the HARPS instrument on the
ESO 3.6 m telescope at La Silla (Chile), under the GTO program ID
072.C-0488 and the regular programs: 085.C-0019, 087.C-0831 and
089.C-0732.
(cid:63)(cid:63) RV data are only available in electronic form at the CDS
tion induced on the parent star by the systems requires a substan-
tial observational effort both in terms of observing time and of
the measurement precision.
time at
Our survey is using the HARPS spectrograph, which is
mounted full
the 3.6m ESO telescope in La Silla
and is optimized for precise radial velocity measurements
(Mayor 2003). To date, it has discovered more than 150 exo-
planets, most of them (through the high-precision program) with
masses in the super-Earth range; this definitely changed our view
on the field.
Since the start of HARPS operations we are monitoring
about 850 stars with moderate radial velocity precision (≈ 2 −
3ms−1), searching for Jupiter and Neptune planets orbiting solar-
type stars (Lo Curto 2010), while the HARPS high-precision
program (Mayor 2003), which routinely achieves precision bet-
ter than 1ms−1, targets lower mass planets, mostly in the super-
Earth mass regime. The goals of our program are the acquisi-
tion of accurate orbital elements of Jupiter and Neptune-mass
planets in the solar neighborhood (within 57.5pc) and the search
for Jupiter twins, which might reveal solar system twins. Of the
43 planets discovered so far by this program, 17 have periods
longer than 1000 days (Moutou 2009, Naef 2010, two more are
presented in a paper in preparation), but their eccentricities are
significantly higher than that of Jupiter. The program has de-
tected three systems with two planets and one triple planetary
system so far (H´ebrard 2010, Lo Curto 2010). Out of the nine
planets in multiple systems we have discovered so far, only two
1
G. Lo Curto et al.: The HARPS search for southern extrasolar planets
have masses larger than or equal to Jupiter, which is consistent
with the observed trend of low-mass planets being detected pri-
marily in multiple systems.
In this paper we report three stars with multiple Keplerian-
like signals. In section 2 we discuss the stars' properties, and in
section 3 the orbital solutions. In section 4 we briefly discuss the
encountered systems.
2. The stars
The three stars belong to our volume-limited sample and
were selected to be low-activity, non-rotating, solar-type stars
on the main sequence (Lo Curto 2010). We have used the
magnitude estimates by Hog (2000), and the one by Koen
(2011) for the redder star BD-061339. They originate from
the Tycho and HIPPARCOS catalogs, respectively (ESA-1997,
Van Leeuwen 2007). The color indices and the parallaxes are
from Van Leeuwen 2007, and the bolometric corrections are ex-
tracted from a calibration to the data presented by Flower (1996),
with uncertainties generally below 0.1 mag.
2.1. HD103774
The star HD103774 is a young, metal-rich, main-sequence star
with spectral type F5V (Pickles 2010, Sousa 2011). Its average
activity level is low, but it varied considerably during the time
span of our observations (see figure 4), going from a more "ac-
tive" state with log(RHK) = −4.72 to a more "quiet" state with
log(RHK) = −5.00 (see Noyes 1984 for a definition of the ac-
tivity indicator). We therefore expect to measure some level of
radial velocity jitter.
Using the online data service VizieR, we found measure-
ments of this star in the mid- and far-infrared, from the IRC (in-
strument onboard the AKARI satellite, Ishihara 2010) and IRAS
(Moshir 1990) catalogs (see figure 1). The data points at 25, 60,
and 100 µm are upper limits. The solid line in figure 1 reproduces
the black body corresponding to the star's photospheric tempera-
ture of 6489K. The uncertainty on the temperature (77K) would
not be noticed on the scale of this graph. The point at 12µm
is deviating from the black body and might seem to indicate
an infrared excess. Circum-stellar dust is indeed quite common
around early-type young stars. While the evidence for infrared
excess at 12 µm is only marginal (1 sigma) and calls for more
precise measurements, no excess is measured up to 9 µm, giv-
ing an indication on the minimum size of the grains that could
possibly be present in the disk.
2.2. HD109271
This old and quiet G5V star does not show infrared excess
(Ishihara 2010, IRAS data are not available), and has a near-
solar metallicity (Sousa 2011). Its basic parameters, as for the
other stars, are shown in table 1.
2.3. BD-061339
BD-061339 is a late-type dwarf (spectral type K7V/M0V), with
no obvious infrared excess (Moshir 1990, IRAS Faint Source
Reject Catalog, Version 2.0).
The estimation of the metallicity in late-type dwarfs is com-
plicated by the abundance of molecular bands of various metal
oxides in the spectra (e.g. TiO, VO). Neves et al. (2012) have
discussed various calibrations present in the literature and pre-
2
Fig. 1. Spectral flux density for HD103774 measured in Jy. The
solid line is a black body with the temperature of the star: 6489K.
The arrows indicate that the three data points are only upper lim-
its.
sented an improved photometric calibration for the metallicity.
We adopted their prescription to estimate the metallicity of BD-
061339, and used the dispersion around their calibration as an
estimate of the uncertainty. This value is well above the mea-
surement uncertainty computed via error propagation from their
formula, and is limited by the quality of the calibration.
index log R(cid:48)
in table 1.
Owing to the red color of the star the estimate of the activity
HK is considered unreliable and no value is included
3. Orbital solutions
We studied the radial velocity (RV) variation of our target stars
via the analysis of the periodograms of the RV data following
the generalized Lomb-Scargle periodogram method outlined by
Zechmeister and Kuerster (2009) and via the use of genetic al-
gorithm as described in the work by Segransan et al. (2011). In
all cases we verified whether the RV was correlated with the bi-
sector span of the cross correlation function (CCF), the activity
indicator RHK (Noyes 1984), and the FHWM of the CCF be-
fore claiming a companion. In addition to this, we also studied
the periodogram of these indicators. When attempting a fit with
more degrees of freedom (e.g., using a model with more planets
or adding a long-term drift to the model), we verified with an
F-test that the new fit brings a significant improvement over the
fit with fewer degrees of freedom. We ran an F-test and obtained
a false-alarm probability (FAP) for which we used a threshold
of 10−4 to identify significant peaks. The orbital elements are
finally displayed in table 2.
3.1. HD103774
We acquired 103 radial velocity measurements for this star in
about 7.5 years. The average uncertainty is ≈ 2.6ms−1, while the
standard deviation of the radial velocity over the entire data set is
≈ 26ms−1, a clear indication of variability. The periodogram of
the radial velocities (Fig. 2) indicates an excess of power at ≈ 5.9
days, while no other significant peak is visible. The one-year and
the two-year aliases (at 5.79d and 5.84d) are also present in the
periodogram, but at half the power of the main signal.
In figure 3 we show the Keplerian fit as a function of phase
together with the data. The orbit is evident, although noisy.
structures
("bumps" at ≈ 11d, 13d, 15d, 470d, 740d), they are not signifi-
Although the periodogram shows various
G. Lo Curto et al.: The HARPS search for southern extrasolar planets
Table 1. Observed and inferred stellar parameters of the three stars of this study. π: parallax, d: distance, MV: absolute magnitude in
V, Bol. corr.: bolometric correction, L: luminosity, Te f f : effective stellar temperature, M∗: stellar mass, log g : gravity at the star's
surface, vsin i: projected rotational velocity of the star, log R(cid:48)
HK): star's rotation period
estimated via the activity indicator.
HK : activity indicator (Noyes 1984), Prot(log R(cid:48)
Parameter
Hipparcos name HIP
Spectral type
V
B − V
π
d
MV
Bol. Corr.
L
Teff
[Fe/H]
M∗
log g
age
vsin i
log R(cid:48)
Prot(log R(cid:48)
[mag]
[mag]
[mas]
[pc]
[mag]
[mag]
[L(cid:12)]
[K]
[dex]
[M(cid:12)]
[cgs]
[Gyr]
[km s−1]
[days]
HK
HK)
HD 103774
58263
F5V
7.12 ± 0.01
0.503 ± 0.014
18.03±0.52
55±2
3.40 ± 0.06
-0.013
3.5 ± 0.3
6489 ± 77
0.28 ± 0.06
1.335±0.03
4.26 ±0.13
1.05 ± 0.64
8.1
-4.85 ± 0.07
7
HD 109271
61300
G5V
8.05 ± 0.01
0.658±0.002
16.21±0.75
62±3
4.1 ± 0.1
-0.097
2.0 ± 0.3
5783 ± 62
0.10 ± 0.05
1.047± 0.024
4.28 ± 0.10
7.3 ± 1.2
2.7
-4.99 ± 0.08
24
BD-061339
27803
K7V / M0V
9.69 ± 0.01
1.321 ± 0.001
49.23±1.65
20±1
8.15 ± 0.07
-0.782
0.095 ± 0.01
4324 ± 100
-0.14±0.17
0.7
4.74 ± 0.02
4.4 ± 4.0
0.57
-
-
(Hog 2000, Koen 2010)
(Van Leeuwen 2007)
(Van Leeuwen 2007)
(Flower 1996)
(Sousa 2011, Ammons 2006)
(Sousa 2011, Neves 2012)
(Girardi 2000)
(Girardi 2000, Sousa 2011)
(Girardi 2000)
(Glebocki 2005, CORAVEL)
(Noyes 1984)
Fig. 2. Periodogram of the RV data for HD103774. A significant
peak is visible at ≈ 5.9 days. The topmost dashed line is the
10−4 false-alarm probability, the lower dotted lines correspond
to 10−3, 10−2, and 10−1 false-alarm probabilities.
Fig. 3. Keplerian fit of the radial velocity data of HD103774 as
a function of orbital phase.
cant, and we see no correlation of the radial velocities with the
activity indicators S MW (Baliunas 1995) or log(RHK) (Pearson
coefficient < 0.1), nor with the bisectors or the FWHM. The
time series of the activity index log(RHK) shows strong vari-
ations of the index with time (Fig. 4), indicating that the star
activity level might increase considerably.
Fig. 4. Variation of the radial velocity (upper panel) and the ac-
tivity index (lower panel) of HD103774 with time.
Because we have enough data points, we separately analyzed
three groups of RV data in three disjoint Julian date bins. From
orbital fitting we have obtained the same periods for the three
groups of data within the experimental uncertainty of ≈ 10−3
days. The same is true for the phase of the orbit, although in
this case the uncertainties are larger (≈ 2 days). Therefore the
detected period is constant within the time span of the measure-
ment, i.e., 7.5 years.
The period of ≈ 5.9 days is quite close to the estimated ro-
tation period of the star (Noyes 1984) of ≈ 7 days. However, the
lack of periodicity of the bisector span, together with the lack
of correlation between bisector and radial velocity, and the fact
that the orbital period is strictly constant along the 7.5 years of
data gives us confidence that the most likely cause of the radial
velocity variation with the period of ≈ 5.9 days is a low-mass
companion.
When looking at the residuals we still notice a significant
scatter of the data. Moreover, the periodogram of the resid-
uals shows an excess of power corresponding to a period of
≈ 750 days, jointly with its one-year alias at ≈ 250 days. We
significantly improved the overall fit to the data by fitting two
3
10.0100.01000.2.20.200.2000.5.50.500.5000. 0 5 10 15 20 25 30Period [days]Normalized power . . . .−0.10.00.10.20.30.40.50.60.70.80.91.01.1−60−40−20 0 20 40HD103774harpsmeanfRV [m/s] . . . .G. Lo Curto et al.: The HARPS search for southern extrasolar planets
Fig. 5. Activity index log(RHK) of HD103774 as a function of
the residuals from the single-Keplerian fit. A strong correlation
is evident. The value of the Pearson correlation coefficient is 0.6.
Keplerians. However, we detected a strong correlation between
the radial velocity residuals to the single-Keplerian fit and the
activity index log(RHK) (Fig. 5) with the value of the Pearson
correlation coefficient which is equal to 0.6, and found a peak in
the periodogram of the activity index corresponding to this pe-
riod. For these reasons we attribute the variation of the residuals
to the one-Keplerian fit to stellar magnetic cycles (Santos 2010).
The highly varying log(RHK) reaches values as high as −4.7,
lending further support to the suggestion that the activity of this
early-type, relatively young star could be responsible for the
large χ2 and the scatter of the residuals around the one-planet
fit.
3.2. HD109271
We measured the radial velocity of the star HD109271 95 times
in 7.5 years. The RV varies by 6ms−1 RMS, well in excess of the
average RV error, which is below 1ms−1. The periodogram of
the observed data is shown in the top panel of figure 6, and two
peaks are clearly visible, one of which is above the 10−4 FAP
level. We fit a Keplerian to this signal and obtained a good fit for
a period of ≈ 7.85 days.
After fitting the first period, looking at the periodogram of
the residuals (figure 6, middle panel) we notice a second peak,
corresponding to the second-highest peak in the periodogram of
the observations, above our FAP limit. This signal corresponds
to a variation with a period of ≈ 30.98 days.
The phased radial velocity curves of the two planets result-
ing from a two-Keplerians (simultaneous) fit and after subtrac-
tion of the RV signal of the other planet are shown in figure 7.
Neither of the signals is correlated with the bisector, with the
activity index, or the FWHM of the CCF; we also verified that
the periodograms of these parameters have no power at periods
corresponding to the identified orbital periods. Moreover, both
signals were identified using the entire data set and also using
three disjoint subgroups of the entire data set, indicating that
they are present during the whole time span of our observations.
Therefore we consider the origin of the two signals to be two
low-mass companions orbiting the star HD109271.
After the two-planets fit to the HD109271 RV data, the pe-
riodogram of the residuals shows two signals above our FAP
threshold at P≈ 198d and P≈ 430d (figure 6, bottom panel).
Within the uncertainties, these periods are the one year alias
4
Fig. 6. Periodograms of HD109271. From top to bottom: peri-
odogram of the data, the residuals after the one-planet fit, and
the residuals after the two-planets fit. The dashed line marks the
10−4 FAP, the dotted lines, from top to bottom, the 10−3, 10−2,
and 10−1 FAP.
of each other. When constraining the analysis to consider time
spans shorter than one year but longer than 200 days, both peaks
disappear, and they reappear together only when the time span
is longer than ≈ 600 days, an indication that the signal could
be the peak at ≈ 430 days, and the one at ≈ 198 days could
be the alias. The third signal, which could be associated to a
1.3 Neptune-mass planet on a 1 AU eccentric (e≈ 0.36) or-
bit, is very noisy and at the limit of detectability by our survey
(K1 < 2ms−1, < δRV >≈ 1ms−1). Although we cannot asso-
ciate this signal to stellar activity, we need a larger data set and
a higher precision to eventually confirm it as a companion to
HD109271.
3.3. BD-061339
We collected 102 radial velocity measurements of the star BD-
061339 in about eight years . The average RV uncertainty is
2.7 ms−1. Figure 8 shows the periodogram of the observations,
clearly indicating an excess of power at ≈ 125 days, well above
the 10−4 FAP level.
A Keplerian fit to the data allowed us to reconstruct an or-
bit whose RV variation as function of the orbital phase is shown
in Figure 9. The orbital fit is very robust, no correlation is seen
between the radial velocities and the activity index, the bisector
span or the FWHM of the CCF, and subsamples of the data se-
lected in Julian date show a periodic RV variation with the same
period within the uncertainties.
-5.00 -4.95 -4.90 -4.85 -4.80 -4.75 -4.70 -40.0 -30.0 -20.0 -10.0 0.0 10.0 20.0 30.0log10(RHK)Residuals from single keplerian fit (m/s)!!10.0100.01000.2.20.200.2000.5.50.500.5000. 0 5 10 15 20 25Period [days]Normalized power . . . .10.0100.01000.2.20.200.2000.5.50.500.5000. 0 5 10 15 20Period [days]Normalized power . . . .10.0100.01000.2.20.200.2000.5.50.500.5000. 0 2 4 6 8 10Period [days]Normalized power . . . .G. Lo Curto et al.: The HARPS search for southern extrasolar planets
Fig. 10. Periodogram of the residuals to the first Keplerian fit to
the of BD-061339 RV data. A signal at P≈ 3.9 days is visible
just above the 10−4 FAP level.
Fig. 7. Radial velocity curves of planets HD109271b (top panel)
and HD109271c (bottom panel) as a function of phase.
Fig. 8. Periodogram of the observations of BD-061339. The pe-
riod at ≈ 125 days is clearly visible well above the 10−4 FAP
level.
Fig. 9. Orbital fit of the 125-day signal to the BD-061339 RV
data.
days with the signal just above the 10−4 FAP level.
The periodogram of the residuals shows an excess at P≈ 3.9
The peak in the periodogram at ≈ 3.9 days is well defined,
and the fit with a two-planets model converges and significantly
improves the χ2. The F-test gives a 99% probability of improve-
ment of a two-Keplerian simultaneous fit over a one-Keplerian.
The uncertainty on the period from the fit is less than 10−3
Fig. 11. Evolution in time of the integrated flux in the Ca II H
line of BD-061339. The star appears much less active during
the last observation campaigns. The red points (darker) represent
the amplitude of the Ca II H re-emission normalized at 6540Å,
while the green points (lighter) represent the S MW index. The
latter one is heavily affected by the low S/N in the blue side of
the spectrum.
days due to the many periods covered through the entire ob-
servations campaign. Because of the quite red color of the star
(B − V = 1.321), the activity indicator log(RHK) is not reliable.
In an attempt to study the activity of this evolved star, we ob-
served the evolution of the re-emission in the Ca II H and K lines
with time. To quantify the re-emission we normalized the spec-
tra to the continuum. As suggested by Mould (1976), the region
around 6540Å is only weakly affected by TiO bands. By direct
inspection we verified that the contribution of the TiO bands at
this wavelength for this star is indeed negligible, and that the
width of the Hα line is such that it does not affect the region we
planned to use to estimate the continuum. The normalized inte-
grated flux of the Ca II H line as function of time is shown in
figure 11. It drops by a factor ≈ 2 between JD 2454000 and JD
2456000. In figure 11 we also show the time dependence of the
S MW index: this index drops as well as a function of time, al-
though the slope is significantly smaller. The S MW index is nor-
malized to the continuum much closer to the re-emission line,
but the signal-to-noise ratio (S/N) is very low in this region,
and the substantial photon noise strongly influences the measure-
ment. For this reason we attempted the normalization at 6540Å.
The star is becoming less active in recent times. Figure 12
shows the re-emission in the Ca II H line in April 2007 (JD =
2454200) and September 2012 (JD = 2456200). The intensity of
the line almost halved during this period.
5
!!−0.10.00.10.20.30.40.50.60.70.80.91.01.1−5 0 5HD109271harpsmeanφRV [m/s] . . . .−0.10.00.10.20.30.40.50.60.70.80.91.01.1−5 0 5HD109271harpsmeanφRV [m/s] . . . .10.0100.01000.2.20.200.2000.5.50.500.5000. 0 5 10 15Period [days]Normalized power . . . .−0.10.00.10.20.30.40.50.60.70.80.91.01.1−10−5 0 5 10BD−061339harpsmeanfRV [m/s] . . . .10.0100.01000.2.20.200.2000.5.50.500.5000. 0 5 10 15Period [days]Normalized power . . . .G. Lo Curto et al.: The HARPS search for southern extrasolar planets
Fig. 12. Ca II H re-emission from two spectra of BD-061339
acquired in April 2007 and September 2012. The intensity of the
line of the more recent spectrum is reduced almost by half.
Fig. 14. Mass-period distribution of the RV-detected planets. The
crosses represent all detected planets, the filled triangles the
planets from this program, the asterisks the solar system plan-
ets, and the filled circles the planets presented in this work.
activity, at other times detection of planets around these stars
may become possible.
Fig. 13. Radial velocity as function of the orbital phase of the
short-period planet of the star BD-061339.
Because of the variation of the activity level of the star with
time, we divided our data set into two subsamples: one contain-
ing data acquired before JD = 2454500 and the other being its
complement. These subsamples contain about the same number
of measurements, therefore our ability to identify the periodic
variation induced by the potential second planet should be the
same in both of them (the sampling is roughly similar). The nor-
malized power at the peak in the periodogram goes from 3.9
(arbitrary units) for observations taken before JD=2454500 to
10.1 for observations taken after this date, the power increases
for each new added data point. The period stays constant within
one sigma in the two datasets. It is worth underlining that the
uncertainty on the period is small also for the two separate sub-
samples: less than 10−2 days.
Adding the fact that we see no correlation between the RV
residuals to the one-planet fit and the FWHM of the CCF or
the bisector span, nor do we see significant power in the peri-
odogram of these parameters around the ≈ 3.9 days period, we
gain confidence that the second signal, which is becoming more
evident as the star becomes less active, is indeed a planet. The
RV as a function of the orbital phase is shown in figure 13.
This case is nice and instructive, and clearly shows how a
planetary signal emerges from the noise of an active star when
the activity level drops. It also reminds us that "active" stars are
not necessarily always active, and that if at some time they are
inappropriate for a planet search because of their high level of
6
4. Discussion
We have identified five planetary signals with m sin(i) going
from 0.6 to 7 times the mass of Neptune. Assuming they lie in the
same region of the radius-mass space as the confirmed transiting
planets (Schneider 2012), we can estimate their radius and there-
fore their density. With this assumption, we obtain a maximum
density for these newly discovered planets that is comparable to
Neptune's or Jupiter's mean density, making them most likely
gaseous planets.
Observing the current detections in the mass-period diagram
(figure 14), we see that the planets presented in this paper are
close to the detection limit of our survey, of about ≈ 3ms−1, and
that two of them fall in the low-populated region of intermediate
periods, between ≈ 10 and ≈ 200 days. This region is indeed
mostly occupied by planets detected in multiple planetary sys-
tems. It makes us wonder whether the presence of more than
one planet breaks a certain symmetry that then allows the planet
to occupy a range of otherwise "forbidden" orbital periods.
The periods of the two Neptune-mass planets orbiting
the star HD109271 are near a 1:4 resonance, but this would
require a thorough dynamical analysis to be confirmed. The
planetary system orbiting the star HD109271 is very similar to
the HD69830 system (Lovis 2006): the periods of the planets
b and c differ only by 10% and 2% respectively, while the
planet masses differ by less than a factor of two, which is an
indication that there seem to be some "preferred architectures"
among planetary systems. Whether this is real or an effect of
our limited understanding of the reality is an open question.
The outer planet of the system orbiting the star BD-061339
is particularly interesting because it falls right in the middle of
the habitable zone (HZ) of its star. Following the method out-
lined by Kane (2012), we estimate the habitable zone of the star
to extend from 0.5 to 0.9 AU. The orbit of planet c, even consid-
ering its relatively high eccentricity of 0.31, lies entirely within
−0.10.00.10.20.30.40.50.60.70.80.91.01.1−10−5 0 5 10 15BD−061339harpsmeanfRV [m/s] . . . . 0.1 1 10 100 1000 10000 100000 1 10 100 1000 10000Period (days)M2 sin (i) (MEarth)Time span limit (9 years)1m/s limit3m/s limit10m/s limit All discovered exoplanets Solar system planets HARPS planets from this program HARPS planets from this workG. Lo Curto et al.: The HARPS search for southern extrasolar planets
Table 2. Orbital parameters for the planets. P: period, T: time at periastron passage, e: eccentricity, ω: argument of periastron, K:
semi-amplitude of the radial velocity curve, m sin i : projected planetary mass, a: semi-major orbital axis, V: average radial velocity,
Nmeas: number of measurements, < δ RV>: average uncertainty on the radial velocity measurements, σ (O-C) : root mean square of
the residuals from the Keplerian fit.
Parameter
P
T
e
ω
K
m sin i
m sin i
a
V
Nmeas
Time span
< δ RV>
σ (O-C)
χ2
red
[days]
[JD-2400000]
[deg]
[m s−1]
[MJup]
[MEarth]
[AU]
[km s−1]
[days]
[ms−1]
[ms−1]
HD 103774b
5.8881 ± 0.0005
55675.4 ± 2.0
0.09 ± 0.04
-42±64
34.3 ± 1.8
0.367 ± 0.022
116 ± 7
0.070 ± 0.001
-3.049±0.003
102
2734
2.6
11.43
15.7 ± 0.6
HD 109271b
7.8543 ± 0.0009
55719.1 ± 3.6
0.25 ± 0.08
-64±20
5.6 ± 0.5
0.054 ± 0.004
17 ± 1
0.079 ± 0.001
-4.903±0.001
87
2683
0.9
2.05
4.6 ± 0.3
HD 109271c
30.93 ± 0.02
55733 ±8
0.15 ±0.09
4± 120
4.9 ± 0.4
0.076 ± 0.007
24 ± 2
0.196 ± 0.003
-4.903± 0.001
87
2683
0.9
2.05
4.6 ± 0.3
BD−061339b
3.8728 ± 0.0004
55220.50 ± 0.10
0.0 (fixed)
0.0 (fixed)
4.4 ± 0.6
0.027 ± 0.004
8.5 ± 1.3
0.0428 ± 0.0007
23.625 ± 0.003
102
2955
2.7
4.3
2.47 ± 0.23
BD−061339c
125.94 ± 0.44
55265.2 ± 17.6
0.31 ± 0.11
41 ± 55
9.1 ±2.9
0.17 ± 0.03
53 ± 8
0.435±0.007
23.625 ± 0.003
102
2955
2.7
4.3
2.47 ± 0.23
the HZ. Depending on the albedo (in the range 0.4 - 0.9) and on
the atmospheric model (efficient / inefficient heat transport) the
surface temperature of the planet can be between 0 oC and 100
oC. Although suggestive, this is a simplistic approach however,
because it does not take into account that the albedo as well as
the stellar flux vary with wavelength. Moreover, due to its mass,
the planet is likely to be a gas planet, making conditions for life
probably more difficult.
rence is generally low. Interestingly, most of the planets with
periods in this range belong to multiple planetary systems.
Owing to the small amplitude of the signals and to the prox-
imity of their amplitudes, approximately 100 data points per sys-
tem were necessary to achieve the current detections, which sup-
ports ample time coverage and high-cadence monitoring as the
key elements to access some of the most interesting planetary
detections.
5. Conclusions
We have presented new results from our nine-year low-precision
(2 − 3ms−1) survey of a volume-limited sample of the solar
neighborhood with HARPS. We detected five exo-planets orbit-
ing three low-activity, slowly rotating solar-type stars. Four of
the detected planets are in two multi-planet systems. The planets
have m sin(i) ranging from 0.6 to 7 Neptune masses, the lightest
being in the super-Earth mass range, underlining that our survey,
albeit limited in precision because of a strategical choice, has the
capability to detect Neptune-mass and in some cases even super-
Earth planets.
The star BD-061339 hosts two planets: in the case that
sin(i) = 1, they would be a super-Earth, possibly in a tidally
locked configuration due to its proximity to the host star, and
a planet of ≈ three Neptune masses in the habitable zone. The
case presented by this star is particularly instructive because it
shows how one of the signals became more clear with time, as
the stellar activity was diminishing.
The star HD103774, hosting one planet with m sin(i) roughly
the mass of Neptune, has also shown signs of varying activity
during the nine years of our monitoring campaign. A reminder
that the stellar activity changes with time, and stars that at one
time are unsuitable for high-precision RV measurements, might
become more quiet at another time and offer the possibility to
acquire precise RV data.
The star HD109271 hosts two planets with m sin(i) approxi-
mately the mass of Neptune and with orbital periods and m sin(i)
very similar to planets b and c around the star HD698430
(Lovis 2006). This fact suggests that some planetary architec-
tures might be "preferred" in Nature.
Two out of the five planets we detected have periods in the
range between 10 and 200 days, a region where planet occur-
Acknowledgements. We thank the staff of the La Silla Observatory at ESO in
Chile for their passionate and professional support.
This research has made use of the VizieR catalogue access tool, CDS,
Strasbourg, France. The original description of the VizieR service was published
in A&AS 143, 23.
NCS
support
acknowledges
European Research
Council/European Community under
the FP7 through Starting Grant
agreement number 239953, as well as from Fundac¸ao para a Ciencia e a
Tecnologia (FCT) through program Ciencia 2007 funded by FCT/MCTES
(Portugal) and POPH/FSE (EC), and in the form of grants reference
PTDC/CTE-AST/098528/2008 and PTDC/CTE-AST/098604/2008.
the
by
the
We thank the referee Martin Kurster for the helpful comments which con-
tributed to improve this paper.
References
Ammons, S.M., et al., 2006, ApJ, 638, 1004A
Baliunas, S.L. et al., 1995, ApJ, 438, 269
ESA, 1997, ESA-SP 1200 "The HIPPARCOS and TYCHO catalogue"
Fischer, D. et al., 2012, ApJ, 745, 21
Flower, P.J., 1996, ApJ, 469, 355
Girardi, L. et al., 2000, A&AS, 141, 371
Glebocki R, et al., Vizier online data catalog, 2005, ESA SP-560, 571
H´ebrard, G. et al., 2010, A&AS, 512, A46
Hog, E. et al., 2000, A&A, 355, L27
Ishihara, D. et al., 2010, A&A, 514A, 1I
Kane, S. et al., 2012, PASP, 124, 914
Koen, C. et al., 2010, MNRAS, 403, 1949K
Lo Curto, G. et al., 2010, A&A, 512,A48
Lovis, C. et al., 2006, Nature, 441, 305
Lovis, C. et al., 2011, A&A, 528, A112
Mayor, M., Queloz, D., 1995, Nature, 378, 355
Mayor, M. et al., 2003, The Messenger, 114, 20
Moshir, M. et al., 1990, IRASF.C, 0M
Mould, J. R., 1976, ApJ, 207, 535
Moutou, C. et al., 2009, A&A, 496,513
Naef, D. et al., 2007, A&A, 470,721
Naef, D. et al., 2010, A&A, 523, A15
Neves, V. et al., 2012, A&A, 538,A25
Noyes, R.W. et al., 1984, ApJ, 279, 763
7
G. Lo Curto et al.: The HARPS search for southern extrasolar planets
Pickles, A., De Pagne, E., 2010, PASP, 122, 898, 1437
Santos, N.C. et al., 2010, A&A, 511, A54
Schneider, J., http://exoplanet.eu, 15 October 2012
Segransan, D. et al., 2011, A&A, 535, A54
Sousa, S. et al., 2011, A&A, 533A, 141S
Van Leeuwen, F. et al., 2007,"Validation of the new Hipparcos reduction", 2007,
A&A, 474, 653
Zechmeister, M., Kuerster, M., 2009, A&A, 496, 577Z
8
|
1602.03204 | 1 | 1602 | 2016-02-09T21:51:50 | Identifying False Alarms in the Kepler Planet Candidate Catalog | [
"astro-ph.EP"
] | We present a new automated method to identify instrumental features masquerading as small, long period planets in the \kepler\ planet candidate catalog. These systematics, mistakenly identified as planet transits, can have a strong impact on occurrence rate calculations because they cluster in a region of parameter space where Kepler's sensitivity to planets is poor. We compare individual transit-like events to a variety of models of real transits and systematic events, and use a Bayesian Information Criterion to evaluate the likelihood that each event is real. We describe our technique and test its performance on simulated data. Results from this technique are incorporated in the \kepler\ Q1-17 DR24 planet candidate catalog of \citet{Coughlin15}. | astro-ph.EP | astro-ph | Draft version February 11, 2016
Preprint typeset using LATEX style emulateapj v. 5/2/11
6
1
0
2
b
e
F
9
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
0
2
3
0
.
2
0
6
1
:
v
i
X
r
a
IDENTIFYING FALSE ALARMS IN THE Kepler PLANET CANDIDATE CATALOG
F. Mullally1, Jeffery L. Coughlin1, Susan E. Thompson1, Jessie Christiansen2, Christopher Burke1,
Bruce D. Clarke1, Michael R. Haas3
Draft version February 11, 2016
ABSTRACT
We present a new automated method to identify instrumental features masquerading as small, long
period planets in the Kepler planet candidate catalog. These systematics, mistakenly identified as
planet transits, can have a strong impact on occurrence rate calculations because they cluster in
a region of parameter space where Kepler's sensitivity to planets is poor. We compare individual
transit-like events to a variety of models of real transits and systematic events, and use a Bayesian
Information Criterion to evaluate the likelihood that each event is real. We describe our technique and
test its performance on simulated data. Results from this technique are incorporated in the Kepler
Q1-17 DR24 planet candidate catalog of Coughlin et al. (2015).
Subject headings: planetary systems, eclipses
1.
INTRODUCTION
their sample (see Figure 14 in Burke et al.).
One of the major goals of the Kepler mission (Koch
et al. 2010) is to estimate the frequency of Earth-size
planets in the habitable zones of solar-type stars. To that
end, the spacecraft collected 4 years of near-continuous
data on ∼150,000 stars, searching for the faint signal of
small transiting planets.
The Q1-Q16 catalog of Kepler planet candidates (Mul-
lally et al. 2015) reported 554 new candidate planets.
They noted an excess of candidates at periods longer
than 100 days over what would be expected if planets
were evenly distributed in orbital period. They iden-
tified two large populations of long-period false alarm
Threshold Crossing Events (TCEs, or the statistically
significant periodic events in a lightcurve that are vetted
to produce the planet candidate catalog). One narrow
peak centered around 372 days (the orbital period of the
spacecraft), and a larger, broader peak spanning 200-
600 days (see their § 3.1). While their vetting process fil-
tered out most of the false alarm events from the narrow
peak (leaving no excess of candidates at this orbital pe-
riod), they concluded the observed excess of long-period
planet candidates was more likely due to incomplete fil-
tering of the broader peak than any super abundance of
small planets in this period range (see their § 9.1).
Burke et al. (2015) computed the occurrence rates of
planets as a function of radius and period around solar
analogs using the catalog of Mullally et al. (2015). They
found a sharp rise in the computed frequency of Earth
analogs for periods longer than 300 days. They traced
the excess to 5 planet candidates with radii < 1.2 R⊕
and periods of 450–550 days, a region of parameter space
where such planets could only be detected around 1% of
[email protected]
1 SETI/NASA Ames Research Center, Moffett Field, CA
94035, USA
2 NASA Exoplanet Science Instititute, California Institute of
Technology, Pasadena, CA 91125
3 NASA Ames Research Center, Moffett Field, CA 94035, USA
It is clear that understanding the reliability of the Ke-
pler catalog (the fraction of candidates actually due to
transiting planets) is an important precondition to mea-
suring terrestrial planet occurrence rates. Much work has
been done identifying astrophysical false positives due to
eclipsing binaries and stellar multiplicity scenarios (e.g.
Torres et al. 2011; Morton 2012; Col´on et al. 2015; San-
terne et al. 2012; Bryson et al. 2013; D´esert et al. 2015),
but less on instrumental artifacts. The statistical val-
idation techniques used on Kepler planets (e.g. Torres
et al. 2011; Rowe et al. 2014) compare the probability
that a transit was due to a planet to various eclipsing
binary type scenarios. However, small, long period Ke-
pler candidates have a non-negligible probability of being
caused by an instrumental or processing artifact. Rowe
et al. (2014) explicitly avoided low signal to noise cases
for this reason.
In this paper we report on a new method to identify
and reject false alarm candidates at long periods. This
method, which we dub Marshall4,5, fits the individual
transit events for candidates with models of transits and
commonly found artifacts, and evaluates which one is
more probable based on a Bayesian Information Crite-
rion. Based on simulations, we find that known artifact
types are rejected ≈ 60–70% of the time, while transits
are preserved at the 95%, level as discussed in § 3.
Marshall is one component of the Kepler Robovetter,
an automated process for vetting planet candidates in
Kepler data, which includes the Flux Robovetter (Cough-
lin et al. 2015), the Centroid Robovetter (F. Mullally
2015 in prep), the ephemeris matching of Coughlin et al.
(2014), and the machine learning technique of (Thomp-
son et al. 2015) that identifies short period variable stars
mistakenly identified as planets. The Robovetter re-
places the manual vetting approach of previous catalogs
4 After a character in the animated series Paw Patrol.
5 Never let a three year old name your algorithms.
2
Mullally et al.
with an automatic, rules-based technique. In addition to
removing some of the inevitable subjectivity of the man-
ual process, the performance of the Robovetter can be
tested against large numbers of simulated transits.
Marshall can be easily applied to any transit search
where short duration signals (such as instrumental ar-
tifacts) are misidentified as transits.
In any matched
filter approach, such as Box Least Squares (Kov´acs
et al. 2002), or the TPS algorithm used by the Ke-
pler pipeline (Seader et al. 2015), non-transit signals
will be occasionally mistaken for transits.
If the false
alarm signals can be modeled, the Marshall approach
can be used separate true transits from the false alarms.
The Marshall algorithm will be useful analyzing data
from the upcoming TESS and PLATO missions. We
make a reference implementation in Python available at
https://sourceforge.net/projects/marshall.
2. METHOD
The key insight of our technique is that long period
candidates have few enough events that there is sufficient
signal-to-noise in an individual event to discriminate be-
tween valid and invalid transit shapes. By looking at
individual events we have access to information about
the transit shape that can be lost in the folded transit
event.
For each Kepler Object of Interest (KOI) in the Q1-
Q17 DR24 catalog (Coughlin et al. 2015) we extract a
snippet of data 1.5 times the reported transit duration ei-
ther side of the midpoint of each individual transit event
(i.e. once per orbital period of the candidate planet). We
use the transit parameters (orbital period, epoch, tran-
sit duration) from Seader et al. (2015). We obtain the
publicly available lightcurves from MAST6 and use the
co-trended lightcurve which corrects for many instrumen-
tal features (available in the PDCSAP FLUX column of the
lightcurve file; see Fraquelli & Thompson 2014 and Smith
et al. 2012). We then fit each event with the following
set of models:
1. A parabola.
2. A parabola with a negative-going step-wise discon-
tinuity (i.e. a step down) at the reported time of
ingress.
3. A parabola with a positive-going step-wise discon-
tinuity at the reported time of egress.
4. A parabola plus a Sudden Pixel Sensitivity
Dropout event (SPSD; an SPSD is typically caused
by a cosmic ray hit on the ccd). We model the
SPSD shape as
(cid:26)
0
−Ae−τ (t−tingress)
:
:
t < tingress
t > tingress
(1)
where A and τ are free parameters, and tingress is
the reported transit ingress time.
6 http://archive.stsci.edu/kepler
Figure 1. An example instrumental feature from Kepler data
(Kepler Id 4575824, alias K06428.01), showing the functional forms
of the models fit to the event, offset vertically for clarity. The
model in each case includes a parabola to describe the shape of the
continuum.
5. A parabola plus a box shaped transit, defined as
tingress < t < tegress
otherwise
(2)
(cid:26) −d :
0
:
where d is a free parameter, while tingress and tegress
are constrained so that the transit mid-point is
fixed.
We include a parabolic term in each model to describe
the continuum flux (i.e the flux we would expect if the
proposed transit were not present). The algorithm used
for co-trending (PDC, Smith et al. 2012) tries to preserve
stellar variability, so the continuum is often not flat even
on the short time scales of interest here.
We show representative examples of the various mod-
els in Figure 1. Models 2-4 represent the most com-
mon kinds of artifacts we see in the data. Model 1 (the
parabola) catches the case where a single strong artifact
triggers a detection of a possible planet, and the other
reported events show no strong signal. A box (Model 5)
is the crudest possible model for the transit shape, but
the second order details of ingress shape or limb darken-
ing are not expected to be visible in the low SNR case of
a single transit of a small planet.
The discontinuities in our models present a chal-
lenge for many optimization algorithms. We chose the
Amoeba, or Nelder-Mead method (Nelder et al. 1965),
which does not require the first derivative of the model
to be well behaved. We found the Amoeba often con-
verged on a local minimum, so we repeated the fit for
each model with a variety of initial conditions, and se-
lected the best fitting result.
We occasionally found transits that lie very close to
gaps in the data. The fit is often poor when some of
the expected data is missing, so we imposed a stringent
Kepler False Alarms
3
requirement that no more than 25% of the expected ca-
dences in the selected snippet be missing or gapped be-
fore running the algorithm on that event.
We select the preferred model using the Bayesian In-
formation Criterion (Schwarz 1978). This metric rewards
models that fit the data with fewer parameters, and is
given by
BIC = −2 lnL + k ln N
(3)
where k is the number of free parameters in the model
and N is the number of data points fit. L is likelihood
of the fit, defined as
L ∝(cid:89)
(cid:34)
(cid:18) yi − f (tiθ)
(cid:19)2(cid:35)
exp
− 1
2
σ
(4)
where yi is the value of the ith data point, and f (tiθ)
is the value of the model given a set of parameters θ
and evaluated at the time of the ith data point. σ is the
uncertainty assigned to each data point, calculated in a
manner that is robust to outliers in the data
√
σ = 1.4826/
2 × median(abs(yi+1 − yi))
(5)
The normalization ensures the computed value of σ is
consistent with the value expected if the values of the
data points were drawn from a Gaussian distribution.
The Gaussian assumption is frequently invalid for Ke-
pler lightcurves, but any mismeasurement of σ affects all
model fits equally.
The model with the lowest value of the BIC is the pre-
ferred model. We define the fit score as the difference
between the BIC value of the transit fit and the best fit-
ting artifact model. Positive scores mean that the arti-
fact model is preferred over the box model. Kepler cata-
logs adopt the principle of "innocent until proven guilty",
erring on the side of including suspected false positives
instead of incorrectly rejecting planet candidates. We
therefore only reject events as false alarms if they have a
score ≥ +10. Kass et al. (1995) argue that a difference in
BIC value of 10 or more means the evidence against the
disfavored model is very strong, However, the choice of
threshold is arbitrary; in the next section we describe
how we tuned our algorithm on simulated transits to
achieve the desired completeness (fraction of valid tran-
sits passed by the algorithm) at the chosen threshold, at
the expense of high reliability (fraction of false alarms
detected).
3. PERFORMANCE
To evaluate the performance of our algorithm we ran
it on the simulated transit set used to evaluate the com-
pleteness of the Kepler planet finding pipeline as de-
scribed by J. L. Christiansen (2016 in prep) and based
on the technique of Christiansen et al. (2015). They in-
jected a range of physically realistic transits for a range of
planet sizes, orbital periods etc. into the pixel level time-
series data of all Kepler target stars using the method
described in Christiansen et al. (2015). This end-to-end
Figure 2. Performance of the algorithm on simulated data. The
thick black line shows the percentage of simulated individual tran-
sit events passed by Marshall as a function of the chosen threshold.
The thin lines show the fraction of simulated artifacts failed at that
same score. The more positive the score, the more the transit model
is favored. We set a threshold of score > +10 (vertical magenta
line) to mark an event as a false alarm.
modeling means that the effects on transit shape of the
lightcurve generation (PA, Twicken et al. 2010), and sys-
tematic removal (PDC, Smith et al. 2012) modules of the
pipeline are properly accounted for in our simulation.
We selected for our study 104 individual simulated
transit events for planets with periods > 200 days and
radii < 5 R⊕ that were recovered by the pipeline. We
tuned our algorithm until > 95% of these individual
events were passed. We then measured the performance
of this tuned algorithm against simulated artifacts.
As J. L. Christiansen (2016 in prep) did not inject ar-
tifact signals in their simulations, we perform our own.
We add an artifact model to the lightcurves before co-
trending, as produced by the PA module of the pipeline
(the SAP FLUX column of the MAST lightcurve file). We
cotrend the simulated signal by fitting and removing the
appropriate Covariant Basis Vectors produced by the
PDC module of the pipeline (Stumpe et al. 2012). These
vectors represent the coarse systematic signals in Kepler
lightcurves. They are available at MAST7, are described
in § 2.3.4 of the Archive Manual, and a tutorial on their
use is given in Kinemuchi et al. (2012). PDC uses ad-
vanced techniques (Smith et al. 2012; Stumpe et al. 2014)
to apply the vectors to the data in an effort to prevent
over-fitting of stellar variability. For our purposes, we
capture the important behavior of PDC with the compu-
tationally simpler direct fits, so we apply those instead.
We run 104 simulations over a range of targets and model
parameters for each of the discontinuity and SPSD mod-
els.
We show our results in Figure 2. The thin blue and
orange lines show our performance at identifying and
rejecting systematic events in our simulations (discon-
7 http://archive.stsci.edu/missions/kepler/cbv
4
Mullally et al.
2. Count Nskip, the number of events where some
data, but less than 75% of the expected cadences,
were collected. These events are not tested, be-
cause the gaps often severely bias the fits leading
to inaccurate results. They frequently contain le-
gitimate events so are assumed to pass by default.
3. If Ngood +Nskip ≥ 3 then the KOI passes, otherwise
it fails. This rule is consistent with the mission
requirement of needing at least three transits to
claim the transit detection.
We apply this test to KOIs in the Q1-Q17 DR24 cata-
log with orbital periods > 150 days. We seed our fits with
the transit parameters (period, epoch, duration etc.)
from Seader et al. (2015) and available at the NASA Exo-
planet Archive (Akeson et al. 2013). Although Marshall
is tuned for low SNR events, and ignores second order
effects on transit shape (such as ingress shape and limb
darkening) we find it correctly identifies high SNR events
as transits. Our treatment of Nskip has a small effect on
our results. From the set of 228 TCEs with periods >
150 days, and not already marked as false positives in
the Q1-Q17 DR24 catalog, only 3 additional TCEs are
marked as false positive if we require Ngood ≥ 3.
We show our results in Figure 3 and Table 1. Of the
228 KOIs tested that were not otherwise failed by the
Robovetter, 20 fail our test of having fewer than three
valid transits (i.e Ngood + Nskip < 3) mostly at periods
> 400 days and radii < 3 R⊕. Visual inspection con-
firms the artifact nature of all but two of these KOIs
(K05805.01 and K02758.01). False positive identifica-
tion in the Q1-Q17 DR24 catalog is made entirely by
rule, so these KOIs are marked as false positive in the
final catalog in Coughlin et al. (2015) even though visual
inspection identifies them as legitimate candidates.
5. DISCUSSION
One limitation of our technique is that we must make
an assumption as to how the lightcurve would look in the
absence of a transit at a given epoch. We choose a simple
model of a parabola fit to a small portion of the out-of-
transit lightcurve that we find works well in practice, but
there are cases of stars that exhibit such rapid variabil-
ity that we have difficulty measuring an accurate contin-
uum. This causes some of our events to be misidentified.
A more sophisticated estimate of the continuum would
likely improve performance, but would have to account
for impulsive (i.e. short duration) spacecraft events, as
well as the variability of the star itself.
For the 4% of injected transits that were rejected, we
inspected the data to understand why they failed. We
show an example in Figure 4. The top panel shows the in-
jected event, while the bottom panel shows the lightcurve
after long term trends were removed by PDC. In this
case, the shape of the transit was deformed to make
it look more like a systematic effect. For stars with-
out rapid variations in the lightcurve, this is the most
common reason why simulated planet candidates were
misidentified. However, attempts to use the lightcurves
without the detrending from PDC had significantly worse
Figure 3. Long period KOIs that pass (open circles) and fail
(filled circles) our test. KOIs that fail are clustered at long period
and small radius, corresponding to low SNR detections. Almost all
are detected with only three transits. Two KOIs detected at low
SNR and large radii are not shown on this plot.
tinuities and SPSDs respectively) as a function of the
threshold value. The vertical magenta line indicates the
adopted threshold value of 10 (i.e the transit model has a
BIC score at least 10 points higher than the most favored
artifact model). At a threshold of 10 we reject 60-70% of
the injected events. The thick black line shows the num-
ber of injected transits that are preserved by the tech-
nique, which is > 96%. These results give us confidence
that we can identify many of the false alarm events in
the real data, while preserving the signal of most of the
real planets.
4. APPLICATION TO THE Q1-Q17 DR24 CATALOG
The Q1-Q17 DR24 catalog of KOIs (Coughlin et al.
2015) incorporates results from Marshall. The KOI cre-
ation process is described in detail in Rowe et al. (2015).
Briefly, a KOI number is assigned to a periodic signal
in the lightcurve of a Kepler target that appears to be
due to the transit or eclipse of an astrophysical body.
KOI numbers are assigned based on a preliminary analy-
sis and are sometimes assigned to other phenomena such
as stellar variability or instrumental artifacts. Further
vetting identifies some more of these artifacts as well as
false positive signals due to eclipsing binaries. Any KOI
is marked as a planet candidate unless there is conclusive
evidence that it is not.
A KOI incorporates at least three events equally spaced
in time, and Marshall deals with individual events. To
apply Marshall to KOIs, we need to choose a disposition
(planet candidate or false positive) based on the com-
bined scores of individual events. We adopt the following
rules:
1. Count Ngood, the number of transit events where
> 75% of the expected cadences in the appropriate
interval were collected, and where the computed
score is < +10.
Kepler False Alarms
5
decides if the shape of that event is more likely a transit
or one of a few known artifacts. We apply the algorithm
to the DR24 planet catalog of Coughlin et al. (2015)
and find it rejects 20 small, long period KOIs otherwise
marked as planet candidates by the Robovetter.
Funding for this Discovery mission is provided by
NASAs Science Mission Directorate. All of the data pre-
sented in this paper were obtained from the Mikulski
Archive for Space Telescopes (MAST). STScI is operated
by the Association of Universities for Research in Astron-
omy, Inc., under NASA contract NAS5-26555. Support
for MAST for non-HST data is provided by the NASA
Office of Space Science via grant NNX13AC07G and by
other grants and contracts. This research has made use of
the NASA Exoplanet Archive, which is operated by the
California Institute of Technology, under contract with
the National Aeronautics and Space Administration un-
der the Exoplanet Exploration Program.
Facilities: Kepler
REFERENCES
Akeson, R. L., Chen, X., Ciardi, D., et al. 2013, PASP, 125,
989
Bryson, S. T., Jenkins, J. M., Gilliland, R. L., et al. 2013,
PASP, 125, 889
Burke, C. J., Christiansen, J. L., Mullally, F., et al. 2015,
ArXiv:1506.04175
Christiansen, J. L. 2016 in prep
Christiansen, J. L., Clarke, B. D., Burke, C. J., et al. 2015,
ArXiv:1507.05097
Col´on, K. D., Morehead, R. C., & Ford, E. B. 2015,
ArXiv:1506.07057
Coughlin, J. L., Mullally, F., Thompson, S. E., et al. 2015,
ArXiv:1512.06149
Coughlin, J. L., Thompson, S. E., Bryson, S. T., et al. 2014,
AJ, 147, 119
D´esert, J.-M., Charbonneau, D., Torres, G., et al. 2015,
ApJ, 804, 59
Fraquelli, D., & Thompson, S. E. 2014, Kepler Archive
Manual (KDMC-10008-005)
http://archive.stsci.edu/kepler/documents.html
Kass, Robert E., & Raftery, Adrian E. 1995, Journal of the
American Statistical Association, 90, 773
Kinemuchi, K., Barclay, T., Fanelli, M., et al. 2012, PASP,
124, 963
Koch, D. G., Borucki, W. J., Basri, G., et al. 2010, ApJ,
713, L79
Kov´acs, G., Zucker, S., & Mazeh, T. 2002, A&A, 391, 369
Morton, T. D. 2012, ApJ, 761, 6
Mullally, F. 2015 in prep
Mullally, F., Coughlin, J. L., Thompson, S. E., et al. 2015,
ApJS, 217, 31
Nelder J., A., & Mead, R. 1965, , 7, 308
Rowe, J. F., Bryson, S. T., Marcy, G. W., et al. 2014, ApJ,
784, 45
Rowe, J. F., Coughlin, J. L., Antoci, V., et al. 2015,
ArXiv:1501.07286
Santerne, A., D´ıaz, R. F., Moutou, C., et al. 2012, A&A,
545, A76
Schwarz, Gideon 1978, Ann. Statist., 6, 461
Seader, S., Jenkins, J. M., Tenenbaum, P., et al. 2015,
ApJS, 217, 18
Smith, J. C., Stumpe, M. C., Van Cleve, J. E., et al. 2012,
PASP, 124, 1000
Figure 4. An example of how PDC can deform the shape of a
transit. The top panel shows the lightcurve of Kepler Id 6520870
with a simulated transit (highlighted by the gray bar). The lower
panel shows this same data after co-trending by PDC against data
from other nearby stars to remove coarse trends.
performance due to the more complicated structure in
the out-of-transit flux.
Although the artifact types we test for create most
of the artifacts we are aware of for small, long period
planet candidates, there are presumably other sources
of systematics not yet accounted for that decrease the
sample of planet candidates still further.
In addition,
our simulations suggest we only find two thirds of our
injected false alarms This suggests as many as 10 more
false alarms of the kind we tested for remain in the Q1-
Q17 DR24 catalog. There are 37 planet candidates with
periods longer than 300 days in the catalog. If, similar
to the identified false alarms, the uncaught systematics
also have periods > 300 days, then we expect no more
than 27 (or 73%) of these long period candidates are
actually transits. This places a rough upper bound on the
reliability of the catalog for small, long period planets.
The Kepler Robovetter emphasizes completeness over
reliability. Ensuring that as many planets as possible are
included as candidates in the catalog is a higher priority
than removing as many false positives as possible. Users
of the catalog who place a higher value on reliability may
prefer to set their rejection threshold at at lower BIC
score, at the expense of a decrease in the completeness
of their sample. To this end we provide the Marshall
score for each KOI tested in Table 1.
Finally, we note that the score given in Table 1 is not
the probability a KOI is due to a transit. Even KOIs
with strong BIC scores may be due to non-planetary sig-
nals, such the highly diluted signal of an eclipsing binary
system with small angular separation from the target.
6. CONCLUSION
We present a new technique to identify systematic sig-
nals in Kepler data masquerading as planet candidates.
The algorithm looks at each individual transit event and
6
Mullally et al.
Table 1
KOIs examined by Marshall
KOI
Kepler Id
TCE
NTrans.
Period (days)
K00998.01
K01099.01
K01788.02
K03549.01
K03674.01
K03681.01
K03709.01
K04959.01
K06254.01
K06295.01
1432214
2853093
2975770
2307206
2446623
2581316
2576107
2987433
1433531
2856960
1
1
2
1
1
1
1
1
1
1
9
9
4
7
8
7
7
4
3
7
161.788000(36)
161.52800(34)
369.0790(26)
204.029000 0
175.066000(66)
217.832000(77)
205.58300(16)
420.92900(10)
567.676(13)
204.30400(66)
Radius (R⊕)
30.140(49)
5.97(76)
2.7(2.4)
44.140000 0
28.6(7.1)
11.410(19)
21.0(2.6)
49.4(1.7)
4.27(46)
13.730(62)
Score
−2.1 × 104
−5.5 × 101
+2.7 × 101
−1.4 × 105
−4.2 × 103
−8.0 × 104
−2.8 × 103
−5.0 × 105
+3.1 × 101
−1.4 × 103
Note. - Summary of planet candidate KOIs from Q1-Q17 DR24 with periods greater than 150 days examined by
Marshall. Kepler Id is the unique identifier of the target star in the Kepler input catalog. TCE indicates the order in
which this KOI was found around the target star by the Kepler pipeline in Q1-Q17. NTrans. refers to the number of
observed transits as measured by the pipeline. The uncertainty in the last two digits of period and radius are given in
parentheses. Period and radius values are taken from Seader et al. (2015) and may differ slightly from the final values
in Coughlin et al. (2015). Score represents the confidence Marshall assigns to a KOI. Negative scores indicate high
confidence that a KOI is due to a transit while positive scores indicate lower confidence. KOIs with scores >10 are
deemed to be false positives. The score given in the table is equal to the score of the 3rd strongest individual event
for the KOI, which indicates how close the KOI is to being marked as a false positive. (This table in available in its
entirety in the on-line supplementary materials. A portion is shown here for guidance regarding its form and content.)
Stumpe, M. C., Smith, J. C., Catanzarite, J. H., et al. 2014,
Torres, G., Fressin, F., Batalha, N. M., et al. 2011, ApJ,
PASP, 126, 100
727, 24
Stumpe, M. C., Smith, J. C., Van Cleve, J. E., et al. 2012,
Twicken, J. D., Clarke, B. D., Bryson, S. T., et al. 2010, in
PASP, 124, 985
Thompson, S. E., Mullally, F., Coughlin, J., et al. 2015,
SPIE Ser. 7740, 23
ApJ, 812, 46
|
1312.4329 | 1 | 1312 | 2013-12-16T12:05:26 | The Ginger-shaped Asteroid 4179 Toutatis: New Observations from a Successful Flyby of Chang'e-2 | [
"astro-ph.EP"
] | On 13 December 2012, Chang'e-2 conducted a successful flyby of the near-Earth asteroid 4179 Toutatis at a closest distance of 770 $\pm$ 120 meters from the asteroid's surface. The highest-resolution image, with a resolution of better than 3 meters, reveals new discoveries on the asteroid, e.g., a giant basin at the big end, a sharply perpendicular silhouette near the neck region, and direct evidence of boulders and regolith, which suggests that Toutatis may bear a rubble-pile structure. Toutatis' maximum physical length and width are (4.75 $\times$ 1.95 km) $\pm$10$\%$, respectively, and the direction of the +$z$ axis is estimated to be (250$\pm$5$^\circ$, 63$\pm$5$^\circ$) with respect to the J2000 ecliptic coordinate system. The bifurcated configuration is indicative of a contact binary origin for Toutatis, which is composed of two lobes (head and body). Chang'e-2 observations have significantly improved our understanding of the characteristics, formation, and evolution of asteroids in general. | astro-ph.EP | astro-ph |
The Ginger-shaped Asteroid 4179 Toutatis: New
Observations from a Successful Flyby of Chang'e-2
Jiangchuan Huang1
,
3∗, Jianghui Ji2∗, Peijian Ye1, Xiaolei Wang4,
Jun Yan5, Linzhi Meng3, Su Wang2, Chunlai Li5,
Yuan Li6, Dong Qiao7, Wei Zhao8, Yuhui Zhao2, Tingxin Zhang1,
Peng Liu8, Yun Jiang2, Wei Rao3, Sheng Li9,
Changning Huang10, Wing-Huen Ip6
,
11, Shoucun Hu2, Menghua Zhu6,
Liangliang Yu2, Yongliao Zou5, Xianglong Tang8, Jianyang Li12,
Haibin Zhao2, Hao Huang3, Xiaojun Jiang5, & Jinming Bai13
1China Academy of Space Technology, Beijing 100094, China.
2Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China.
3Institute of Space System Engineering, Beijing 100094, China.
4Beijing Institute of Control Engineering, Beijing 100190, China.
5National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China.
6Space Science Institute, Macau University of Science and Technology, Taipa, Macau.
7Beijing Institute of Technology, Beijing 100081, China.
8Harbin Institute of Technology, Harbin 150001, China.
9School of Electronic Engineering and Computer Science, Peking University, Beijing 100871, China.
10Beijing Institute Of Space Mechanics and Electricity, Beijing 100076, China.
11Institute of Astronomy, National Central University, Taoyuan, Taiwan.
12Planetary Science Institute, AZ 85719, USA.
13Yunnan Astronomical Observatory, Chinese Academy of Sciences, Kunming 650011, China.
∗ The authors contribute equally to this work. To whom correspondence should be addressed. E-mail:[email protected]
1
On 13 December 2012, Chang'e-2 conducted a successful flyby of the near-
Earth asteroid 4179 Toutatis at a closest distance of 770 ± 120 meters from the
asteroid's surface. The highest-resolution image, with a resolution of better
than 3 meters, reveals new discoveries on the asteroid, e.g., a giant basin at the
big end, a sharply perpendicular silhouette near the neck region, and direct
evidence of boulders and regolith, which suggests that Toutatis may bear a
rubble-pile structure. Toutatis' maximum physical length and width are (4.75
× 1.95 km) ±10%, respectively, and the direction of the +z axis is estimated
to be (250±5◦, 63±5◦) with respect to the J2000 ecliptic coordinate system.
The bifurcated configuration is indicative of a contact binary origin for Tou-
tatis, which is composed of two lobes (head and body). Chang'e-2 observations
have significantly improved our understanding of the characteristics, forma-
tion, and evolution of asteroids in general.
Introduction
Chang'e-2, the second Chinese probe dedicated to the exploration of the Moon, was launched
on 1 October 2010 at 10:59:57 UTC. After it accomplished its primary objective, an extended
mission was designed for Chang'e-2 to travel to the Sun-Earth Lagrangian point (L2) and then
to an asteroid. On 9 June 2011, Chang'e-2 departed its lunar orbit and headed directly to
L2, where it arrived on 25 August 2011 and began to perform space-environment exploration.
The asteroid's subsequent flyby mission remained under consideration, and eventually, asteroid
4179 Toutatis was selected as the primary target for the flyby after considering the leftover
fuel of the spacecraft, the capability of the tracking and control network, and the fact that the
asteroid's closest approach to Earth would occur on 12 December 2012. An undisclosed story
is that Chang'e-2 was intentionally controlled to move near L2 for more than 230 days to favor
2
the Toutatis-flyby injection maneuvers. On 15 April 2012, Chang'e-2 left L2. On 1 June 2012,
the spacecraft began its mission to Toutatis. Chang'e-2 implemented the flyby on 13 December
2012 at 8:29:58.7 UTC at a closest distance of 770 ± 120 (3σ) meters from Toutatis' surface
at a high relative velocity of 10.73 km s−1, with highest-resolution images of better than 3
meters per pixel1 . As of 14 July 2013, Chang'e-2 is still traveling into outer space over 50
million kilometers away from Earth, and it is the first successful multiple-type-objective probe
in history that has ever visited the Moon, a Lagrangian point, and an asteroid.
Asteroids are remnant building blocks of the formation of the Solar System. They provide
key clues to understanding the process of planet formation, the environment of the early Solar
nebula, and even the occurrence of life on Earth. Toutatis, as an Apollo near-Earth asteroid in
an eccentric orbit that originates from the main belt, was a good target for Chang'e-2's flyby
mission because this asteroid is full of fascinating puzzles. The 3-D shape of Toutatis has been
well reconstructed, and its spin state has been determined, from extensive radar measurements
taken at Arecibo and Goldstone2 -- 9 since 1992 during its closest approaches to Earth, once every
four years. The results indicate that Toutatis is characterized as a non-principal-axis rotator;
this tumbling asteroid rotates about its long axis with a period of 5.4 days, and the long-axis
precession period is 7.4 days7 ; this precession may be induced by the Earth's tides during each
close-in flyby10 and the Yarkovsky-O'Keefe-Radzievskii-Paddack (YORP) effect11 . Moreover,
detailed geological features, such as complex linear structures and concavities, have been re-
vealed in Hudson et al.'s model5 at an average spatial resolution of ∼ 34 m. Hence, the primary
issue of concern is whether the derived radar models accurately represent the real appearance
of Toutatis. Furthermore, some features such as boulders and regolith on its surface have been
speculated to exist in previous radar models5, 6 . It was hoped that this mystery could be solved
by obtaining optical images during a spacecraft's close flyby. Furthermore, small asteroids at
the km scale are believed to have a so-called rubble-pile structure, which means that they consist
3
of a loose collection of fragments under the influence of gravity. Therefore, the second major
question arises -- whether Toutatis is also an agglomeration of gravitationally bound chunks
with such a fragile structure.
The optical images of Toutatis were obtained using one of the onboard engineering cameras,
which were originally designed to monitor the deployment of solar panels on the spacecraft and
to photograph other objects (such as the Earth and Moon) in space. The camera that is capable
of color imaging has a lens with a focal length of 54 mm and a 1024-by-1024-pixel CMOS
detector, and it has a field of view (FOV) of 7.2◦ by 7.2◦. A solar panel shields the camera from
sunlight to help minimize stray light. The FOV was partly occulted by the solar panel, and thus
some of the images were cut off from the early gallery. As part of the imaging strategy, Chang'e-
2 maneuvered to adjust the camera's optical axis to lie antiparallel to the direction of the relative
velocity vector between the probe and Toutatis. Imaging during the inbound trajectory was not
ideal because there was a large Sun-Toutatis-Chang'e-2 phase angle of ∼ 143.6◦. Therefore,
Chang'e-2 completed an attitude adjustment approximately an hour ahead of the flyby epoch
to prepare for the outbound imaging. The overall imaging time was approximately 25 minutes,
and the interval for each snapshot was 0.2 seconds. During the mission, the spacecraft collected
more than 400 useful images totaling 4.5 GB of data. Figure 1 shows several outbound images
of Toutatis. Panel e, which was the first panoramic image, was taken at a distance of ∼ 67.7 km
at a resolution of ∼ 8.30 m (see Supplementary Figure 1), whereas panel a, which was the third
in a sequence of images, corresponds to an imaging distance of ∼ 18.3 km at a local resolution
of ∼ 2.25 m, as shown in Fig. 1. The present highest-resolution optical image has a resolution
that is significantly finer than the 3.75 m resolution of the known radar model8, 9 , although the
radar images cover nearly the entire surface of Toutatis, whereas the optical images cover only
one side of Toutatis' surface.
4
Results
We tested several methods to directly establish the 3-D model of Toutatis based on the opti-
cal images collected by Chang'e-2. However, there is no satisfactory outcome because of the
narrow viewing angle and the local high-resolution images. Therefore, 3-D models developed
based on delay-Doppler radar measurements5, 8, 9 were adopted to discern the attitude of Toutatis
in the approaching epoch.
Based on the attitude of the probe and its camera, the relative position between the asteroid
and the probe, and optical images, in combination with the models of Hudson et al. and Busch
et al.5, 8, 9 , which were utilized to match the spin state of Toutatis (see Supplementary Figure
2), the direction of the principal axis was measured to be (250 ± 5◦, 63 ± 5◦) with respect to
the J2000 ecliptic coordinate system5, 12 , and the fundamental parameters are summarized in
Table 1. Our result is in good agreement with the previous prediction8 , which indicates that
Toutatis' spin status was greatly changed because of tidal interaction with the Earth during the
2004 flyby, which had a closest distance of ∼ 0.01 AU8, 10 . Furthermore, we again measured
Toutatis' physical size from the images, and its maximum length and width were estimated to
be (4.75 × 1.95 km) ±10%, respectively (see Supplementary Section 3).
New Geological Features from Chang'e-2 Observation
The shape of Toutatis resembles that of a ginger root (Fig. 1), where the bifurcated asteroid
mainly consists of a head (small lobe) and a body (large lobe). The two major parts are not
round in shape, and their surfaces have a number of large facets. In particular, at the end of
the body, there appears to be a huge square concavity, which could be of large-impact origin.
Regarding the asteroid's appearance, craters are more prominent on Toutatis than on Itokawa.
In contrast, we clearly see fewer boulders on Toutatis, although many smaller blocks may have
been missed because of the resolution of Chang'e-2 flyby images.
5
In comparison with the radar models3 -- 5 , the proximate observations from Chang'e-2's flyby
have revealed several remarkable discoveries concerning Toutatis, among which the presence
of the giant basin at the big end appears to be one of the most compelling geological features,
and the sharply perpendicular silhouette in the neck region that connects the head and body is
also quite novel. A large number of boulders and several short linear structures (Fig. 1) are also
apparent on the surface, although the lack of sharp contact in the radar model may be an effect
of the SHAPE software13 .
Concavities
The giant basin at the big end of Toutatis has a diameter of ∼ 805 m, and we surmise that one
or more impactors may have collided with the asteroid in this region, thereby gouging out such
an enormous basin. Judging by the contrast of its relatively subdued relief with large impact
basins on other asteroids, Mathilde and Eros, this large-end depression may be attributable to
the internal structure of the large lobe, with smaller impact features overlying it. The most sig-
nificant feature is the ridge around the largest basin. The wall of this basin exhibits a relatively
high density of lineaments, some of which seem to be concentric to the basin. There are also
several linear structures outside the basin that are roughly parallel to the basin rim (Fig. 2a).
These ridges are indicative of an internal structure on small bodies14, 15 . Most of the ridges near
the largest basin at the big end are most likely related to the huge stress energy during impact.
The interior material in the deep region would have been scattered across Toutatis' surface by
the large impact that formed the basin, leading to the fracturing of the asteroid. Hence, one may
speculate that the asteroid most likely has a rubble-pile internal structure.
More than fifty craters have been identified from the flyby images, and they have diameters
ranging from 36 to 532 m (with an average of ∼ 141 m). ∼ 44% of the craters exceed 100 m
in diameter, which is consistent with Hudson et al.'s model. From the close-up image, a crater
6
of high morphological integrity (with a diameter of ∼ 62 m) can be clearly observed, whereas
tens of boulders (with lengths of ∼ 20-30 m) are randomly distributed nearby (Fig. 2b). Two
concavities are closely carved into the surface: the larger (with a diameter of ∼ 368 m) has a
relatively vague outline, whereas the smaller (with a diameter of ∼ 259 m) has a relatively clear
edge. Based on their appearance, we suspect that the relatively smaller, young crater (marked
with the letter B) is superimposed on the larger, ancient crater (marked with the letter A) (Fig.
2a).
Although the images show that parts of the observed surface areas have a high degree of
exposure, the topographical features of the craters were identified based on visual inspection and
automatic annotation. Therefore, we identified craters on the large and small lobes individually,
including vaguely shaped candidate craters on the highly exposed areas, for the investigation
of the cumulative and relative size-frequency distributions (see Supplementary Section 4)16 .
Figure 3 indicates that the surfaces of the two lobes are very likely to share a similar cratering
history because their relative size-frequency profiles for craters17 look alike throughout a range
of diameters. Furthermore, the craters on the two lobes are not in saturation equilibrium, as
the two profiles do not tend toward stability with increasing crater diameter (Fig. 3a). An
abundance of craters would suggest that Toutatis suffered from many impacts of interplanetary
projectiles in the past.
Boulders
From the flyby images, more than thirty boulders can be clearly discerned. They have lengths
that range from 12 to 81 m, with an average of ∼ 32 m. Approximately 90% of the boulders are
less than 50 m in diameter. The largest boulder, along with other large boulders, is located near
the neck region (Fig. 2a). It is believed that a large ejected fragment is empirically associated
with a large crater18, 19 . The boulders are suggested to be those non-escaping ejecta that re-
7
impact onto the asteroid's surface within a short timescale20 . Recently, the Goldstone radar
images of 12 December 2012 have also suggested that several 10-m-scale bright features may
be boulders8 . However, the optical images, although they were collected on only one side of
the asteroid, have an advantage over the radar models5, 8 in uncovering local fine features on the
surface of Toutatis.
Regolith
Two craters in the neck area are indicative of the granular flows in the inner flanks. The upper
regolith near the craters is fine grained, whereas the lower is relatively coarse, and very many
fragments are present (Fig. 2a). Such a structure may indicate the redistribution of regolith via
downslope or resurfacing processes, which may be related to the inclined terrain and the orienta-
tion of gravity20 . The results are in good agreement with those of polarimetric observations21, 22
, thereby indicating that Toutatis is covered with fine regolith made up of light-transmitting
materials. The same features have also been reported on Lutetia's surface23 . Furthermore,
the characteristics once again suggest that Toutatis' surface may consist of a fine granular or
regolith layer with the porosity of lunar soils4 . The latest results imply that this asteroid may
bear an undifferentiated L-chondrite surface composition24 .
There is additional evidence that supports the existence of regolith on Toutatis' surface. The
average thermal inertia for Toutatis may have a low profile, according to an empirical relation
between the thermal inertia and the effective diameter25 . For a non-principal-axis rotator such
as Toutatis, the global redistribution of regolith that results from lofting may also be one of the
factors that contributes to the low surface thermal inertia25 . A considerable number of mm-
sized grains may constitute Toutatis' regolith (Fig. 2b), and the regolith depth on the surface
might range from several centimeters to several meters. Clearly, Toutatis' thermal inertia seems
to be much less than that of Itokawa26 ; thus, the terrain of Toutatis is a bit smoother than that
8
of Itokawa, and the boulders on Toutatis' surface are less numerous than those on Itokawa's
surface, which is mostly embedded with numerous pebbles and gravel in all regions.
Linear Structures
Several types of linear structures, which are primarily composed of troughs and ridges that
appear to be common for small asteroids27, 28 , can be easily observed on Toutatis' surface.
Notable linear structures are apparent in the radar models5 ; however, the flyby images do not
identify any linear features that have been previously reported because the relevant part of the
surface was invisible to Chang'e-2. The present linear structures, which are similar to the short
linear structures of Itokawa, are clearly observed29 to have a length of 120-330 m in the visible
regions of the images. The troughs, which are similar in appearance to groove-like features, are
primarily scattered over the small lobe; they have an average length of ∼ 170 m, and they are
approximately orientated along the long axis of the asteroid (Fig. 2a). Generally, troughs are
related to extensions resulting from tensile stress in the plane of the surface15 . Consequently,
they may be produced by the tensile stresses that arise from nearby impact cratering or other
geological processes.
Discussion
As mentioned above, the existence of the giant basin and its surroundings provides direct ev-
idence that Toutatis is likely to be a rubble-pile asteroid. In this case, the asteroid could re-
assemble itself into a weak aggregate of large fragments via a heavy impact or many smaller
impacts; in this manner, the large interior voids could absorb the collision energy and further
resist huge collisions in the formation process. Furthermore, a vast majority of S-type asteroids
appear to possess rubble-pile structures with an average porosity of ∼ 15% - 25%30 . Recent
investigations have demonstrated that the typical bulk density of L ordinary chondrites is ∼ 3.34
9
g cm−3 24, 30 . Assuming a density of 2.1 - 2.5 g cm−3 20 , Toutatis may have a porosity in the
range ∼ 25% to ∼ 37%, thereby implying that it may be an intermediate body between Eros
(∼ 20%30) and Itokawa (∼ 41%26) with respect to surface features. Moreover, strong evidence
from N-body simulations indicates that two km-sized objects with rubble-pile structures would
produce Toutatis-like objects31 . In summary, we may conclude that Toutatis is not a monolith
but most likely a coalescence of shattered fragments.
The bifurcated configuration means that Toutatis is comprised of two major lobes, similar
to Itokawa, thereby implying that Toutatis is a contact binary, which is an asteroid type that
may constitute more than 9% of the NEA population. Km-sized contact binaries, such as 4486
Mithra and 4769 Castalia, share an irregular, significantly bifurcated shape32 . The following
question then arises: how do these contact binaries form?
Several formation mechanisms have been proposed to produce such bifurcated configura-
tions. The first scenario supposes that the two major lobes were two separate objects, which
were moving at a sufficiently low relative speed to yield a contact binary26 . The difference in
the orientations of head and body may be indicative of their diverse origins and further suggests
that the two parts were once detached. The second scenario includes the YORP and binary
YORP (BYORP) effects, which are likely to form contact binaries33 , especially for those ob-
jects with slow rotational periods.
In this scenario, two components, which may be major
fragments of a parent body, suffer re-impact and recombination, thereby leading to the forma-
tion of a bifurcated configuration. The third scenario suggests that catastrophic collisions may
have gouged out the giant basin and a number of craters on Toutatis' surface. The boulders near
three potential arc depression terrains (Figure 2a) imply that the head may have experienced
severe impacts. Last but not least, tidal disruption from a close encounter with a terrestrial
planet is also likely to play a role in the formation of such a contact binary. In short, the above-
mentioned scenarios may have actually occurred during Toutatis' history. Hence, we infer that
10
Toutatis may be a reconstituted contact binary.
The Chang'e-2 flyby has revealed several new discoveries regarding Toutatis, e.g., the giant
basin at the big end, the sharply perpendicular profile about the neck region, and direct images of
distributed boulders and surface regolith. It is the first time that geological features on Toutatis'
surface have been observed so clearly. All these observations suggest that Toutatis may bear a
rubble-pile structure, similar to other S-type asteroids, such as Itokawa. Moreover, the close-up
imaging results for the asteroid provide further indication that the bifurcated configuration that
is related to the formation of contact binaries appears to be quite common among the NEA
population. The observations of Chang'e-2 have significantly improved our understanding of
the formation and evolution of these building blocks of the Solar System.
Methods
To adapt to Chang'e-2's characteristics of high control accuracy and the wide FOV of 7.2◦ by
7.2◦ of the engineering camera, a close-approach strategy and a specific imaging design were
developed. As shown in Supplementary Figure 1, L, S, and ∆t represent the minimum distance
of rendezvous, the imaging distance, and the time interval between two sequential images,
respectively. The camera optical axis pointed in the direction of the relative velocity between
the probe and Toutatis, which remained nearly unchanged during the flyby. After a short time
at the closest approach, the camera captured Toutatis; the asteroid was contained in the FOV,
but the imaging size decreased as the departure distance increased. By solving the geometric
relation between Image 2 and Image 3, the distances L and S and the resolution of each image
were determined for the Chang'e-2 observations.
Taking into account the various characteristics of the specific approach design and imaging
strategy for the flyby, a physical size of (4.75 × 1.95 km) ±10% for the illuminated region of
the asteroid surface was statistically estimated from the sequence of images recorded by the
11
CMOS detector during the flyby. The determined size is in good agreement with that of the
radar model5 (see Supplementary Sections 1 and 3).
The attitude of a rigid body in space is determined by its rotations around three axes of an
orthogonal coordinate system. To define the attitude of Toutatis in each optical image (Supple-
mentary Figure 2a), the three axes l1, l2, and l3 are defined as follows: l1 and l2 extend through
Toutatis' centroid in the images along the directions of the long axis and short axis, respectively,
and are perpendicular to each other. l3 is perpendicular to the imaging plane and constitutes a
right-handed coordinate system with l1 and l2.
To match its shape to the optical images, we rotated the model of Busch et al. in intervals
of 5◦ for each Euler angle around the three axes of the body-fixed coordinate system in space.
Next, we chose three criteria of the orientation of the asteroid's long-axis, the ratio of length
to width, and the obvious topographical feature at the joint of the two lobes of Toutatis --
to ensure that the radar model agreed with the optical images from Chang'e-2, as shown in
Supplementary Figure 2a. The most approximate attitude was finally obtained by rotating the
radar model and comparing it with the optical image, as shown in Supplementary Figure 2b.
The coordinate transformation from the asteroid's body-fixed system to the J2000.0 equatorial
coordinate system were expressed in terms of three components of Euler angles. The directions
of the principal axes were then obtained according to the attitude of the radar model.
The craters and other geological features were identified via both visual inspection and auto-
matic annotation. In this case, dozens of low-lying regions were noted, and their positions, con-
tours, and sizes were obtained using surface-structure-analysis techniques. The feature points
on the edge of each crater were labeled manually based on the treatment of the imaging process
and the radar model. After the points were marked, a piecewise-quadratic curve was fitted to
them, and closed curves were obtained as contour profile. The sizes of the craters (and boulders)
were calculated based on the resolution for each image and the fitted contour profiles.
12
References
[1] Huang, J. C. et al. The engineering parameters analysis of (4179) Toutatis flyby mission of
Chang'e-2. Science China Technological Sciences 43, 596-601 (2013).
[2] Hudson, R. S. & Ostro, S. J. Shape and non-principal axis spin state of Asteroid 4179
Toutatis.Science 270, 84-86 (1995).
[3] Ostro, S. J. et al. Radar images of Asteroid 4179 Toutatis. Science 270, 80-83 (1995).
[4] Ostro, S. J. et al. Asteroid 4179 Toutatis: 1996 radar observations. Icarus 137, 122-139
(1999).
[5] Hudson, R. S., Ostro, S. J. & Scheeres, D. J. High-resolution model of Asteroid 4179
Toutatis. Icarus 161, 346-355 (2003).
[6] Busch, M. W. et al. Determining asteroid spin states using radar speckles. Icarus 209, 535-
541 (2010).
[7] Busch, M. W. et al. Twenty years of Toutatis. EPSC-DPS2011 6, 297 (2011).
[8] Busch, M. W. et al. Internal structure of 4179 Toutatis. 2012 AGU Fall Meeting, P31A-1873
(2012).
[9] Takahashi, Y., Busch, M. W. & Scheeres, D. J. Spin State and Moment of Inertia Charac-
terization of 4179 Toutatis, AJ 146, 95 (2013).
[10] Scheeres, D. J. Rotational fission of contact binary asteroids. Icarus 189, 370-385 (2007).
[11] Bottke, W. F., Vokrouhlicky, D., Rubincam, D. P. & Broz, M. in Asteroids III (eds. Bottke,
W. F. et al.)(Univ. of Arizona Press, Tucson, 2002).
13
[12] Scheeres D. J., Ostro S. J., Werner R A, et al. Effects of gravitational interactions on
asteroid spin states. Icarus 147, 106-118 (2000).
[13] Magri, C. et al. Radar observations and a physical model of Asteroid 1580 Betulia Icaurs
186, 152-177 (2007)
[14] Prockter, L. et al. Surface expressions of structural features on Eros. Icarus 155, 75-93
(2002).
[15] Thomas, N. et al. The geomorphology of (21) Lutetia: Results from the OSIRIS imaging
system onboard ESA's Rosetta spacecraft. Planet. Space Sci. 66, 96-124 (2012).
[16] Zou, X.D. et al. Preliminary analysis of the 4179 Toutatis snapshots of the Chang'e-2
fly-by. Icarus, in press (2013).
[17] Arvidson, R. E. et al. Standard techniques for presentation and analysis of crater size-
frequency data. Icarus 37, 467-474 (1979).
[18] Gault, D. E., Shoemaker, E. M. & Moore, H. J. Spray ejected from the lunar surface by
meteoroid impact. NASA Technical Document TND-1767, 1-39 (1963).
[19] Thomas, P. C., Veverka, J., Robinson, M. S. & Murchie, S. Shoemaker crater as the source
of most ejecta blocks on the asteroid 433 Eros. Nature 413, 394-396 (2001).
[20] Scheeres, D. J., Ostro, S. J., Hudson, R. S., DeJong, E. M. & Suzuki, S. Dynamics of
orbits close to Asteroid 4179 Toutatis. Icarus 132, 53-79 (1998).
[21] Mukai, T. et al. Polarimetric observations of 4179 Toutatis in 1992/1993. Icarus 127, 452-
460 (1997).
14
[22] Hudson, R. S. & Ostro, S. J. Photometric properties of Asteroid 4179 Toutatis from
lightcurves and a radar-derived physical model. Icarus 135, 451-457 (1998).
[23] Vincent, J. B., Besse, S., Marchi, S., Sierks, H. & Massironi, M. Physical properties of
craters on asteroid (21) Lutetia. Planet. Space Sci. 66, 79-86 (2012).
[24] Reddy, V. et al. Composition of near-Earth Asteroid (4179) Toutatis. Icarus 221, 1177-
1179 (2012).
[25] Delbo', M., dell'Oro, A., Harris, A. W., Mottola, S. & Mueller, M. Thermal inertia of
near-Earth asteroids and implications for the magnitude of the Yarkovsky effect. Icarus 190,
236-249 (2007).
[26] Fujiwara, A. et al. The rubble-pile asteroid Itokawa as observed by Hayabusa. Science
312, 1330-1334 (2006).
[27] Thomas, P. C. & Prockter, L. M. in Planetary Tectonics(eds. Watters, T. R & Schultz, R.
A.) (Cambridge University Press, New York, 2010).
[28] Robinson, M. S., Thomas, P. C., Veverka, J., Murchie, S. L. & Wilcox, B. B. The geology
of 433 Eros. Meteorit. Planet. Sci. 37, 1651-1684 (2002).
[29] Demura, H. et al. Pole and global shape of 25143 Itokawa. Science 312, 1347-1349 (2006).
[30] Britt, D. T., Yeomans, D., Housen, K. & Consolmagno, G. in Asteroids III(eds. Bottke, W.
F., Cellino, A., Paolicchi, P., & Binzel, R. P.)(Univ. of Arizona Press, Tucson, 2002).
[31] Richardson, D. C., Leinhardt, Z. M., Melosh, H. J., Bottke, W. F. Jr. & Asphaug, E. in
Asteroids III(eds. Bottke, W. F., Cellino, A., Paolicchi, P., & Binzel, R. P.)(Univ. of Arizona
Press, Tucson, 2002).
15
[32] Benner, L. A. M. et al. Near-Earth Asteroid 2005 CR37: radar images and photometry of
a candidate contact binary. Icarus 182, 474-481 (2005).
[33] Steinberg, E. & Sari, R. Binary YORP effect and evolution of binary asteroids. AJ 141, 55
(2011).
[34] Krivova, N. V., Yagudina, E. I, & Shor, V. A. The orbit determination of (4179) Toutatis
from optical and radar data. Planet. Space Sci. 42, 741-745 (1994).
[35] Milani, A., & Gronchi, G. Theory of orbit determination. (Cambridge University Press,
Cambridge, UK, 2010).
Supplementary Information is available in the online version of the paper.
Acknowledgements The authors thank Michael W. Busch for his constructive comments that
greatly helped to improve the original content of this manuscript. We thank the whole Chang'e-
2 team to make the mission a success. We appreciate B. Yang for target observations before the
flyby mission. We are grateful to R.P. Butler for English improvement of this manuscript. This
research was supported by National Science and Technology Major Project, National Natural
Science Foundation of China, the Innovative and Interdisciplinary program by CAS, and Key
Laboratory of Planetary Sciences (2013DP173302), CAS. JYL's contribution to this work is not
supported by any funds.
16
Authors Contributions J. C. H. was the leader of the asteroid mission. J. H. J. wrote the
manuscript with contributions from other co-authors. J.C.H., P.J.Y. and T.X.Z. were responsible
for the overall project design of the flyby mission to Toutatis, imaging strategy, and proposed the
method of fundamental parameter analysis. H.B.Z., J.H.J., S.C.H., S.W., Y.H.Z., Y.J. and L.L.Y.
contributed to ground-based observation, refined the trajectory solution for Toutatis, completed
data analysis, prepared for figures and wrote the manuscript. L.Z.M, H.H. and W.R. took charge
of overall design of probe system and data analysis. X.L.W. was responsible for controlling the
attitude and the spacecraft's orbit, and contributed to the engineering parameters analysis. J.Y.,
C.L.L., Y.L.Z. and X.J.J. took charge of data collection, data pretreatment and the ground-
based observation. Y.L. and M.H.Z. partly contributed to contents and prepared for figure.
D.Q. contributed to the target selection and the design of transfer orbit. W.Z., P.L. and X.L.T.
were in charge of the algorithm of optical images and obtained the image information. S.L.
was responsible for imaging process. C.N.H. developed the optical camera and determined the
parameters of the camera. W.H.I and J.Y.L. contributed ideas and discussions in preparation for
the manuscript. J.M.B. partly participated in the target observation. All authors contributed to
the manuscript.
Authors information The authors declare no competing financial interests. To whom corre-
spondence should be addressed. E-mail:[email protected].
17
Figure 1 Outbound images of Toutatis acquired on 13 December, 2012 during Chang'e-2
flyby, indicative of the spacecraft being away from the asteroid (from a to e). The left side
of Toutatis is blocked by the solar panel in images a-d.The imaging distance (D), epoch of flyby
(T, UTC) and resolution of each image (R) are shown for each snapshot, where the distance
error for a is 1.1 km. The resolution of each image is linear with the distance.
.
18
Figure 2 All kinds of geological features on the Toutatis' surface.
(a) All craters (blue
profiles) and boulders (red squares) are outlined in the panoramic image of Toutatis. Two craters
are closely distributed: the smaller crater (B) seemed to superimpose on the larger one (A).
Green lines indicate the lineaments. Black arrows point the flow direction of the fine-grained
regolith. (b) The enlarged portion shown (white box) in the left panel (a). A morphological-
integrity crater can be observed. Tens of boulders are randomly distributed around.
19
Figure 3 (a)The relative size-frequency distribution for craters on Toutatis. (b) The cumu-
lative size-frequency distribution for craters on Toutatis. In panel (a) R plot was devised by
the Crater Analysis Techniques Working Group17 to better show the size distribution of craters
and crater number densities for determining relative ages. The vertical position of the curve
is a measure of crater density or relative age on the Toutatis: the higher the vertical position,
the higher the crater density and the older the surface. In panel (b) x-axis stands for the crater
diameter and y-axis represents the crater number larger than corresponding diameter on the
investigated area. CSFD would easily inform that the size of craters distribute as diameters
increase.
20
Table 1. Parameters for Toutatis.
Property
Osculating Orbital elements a=2.5336 AU, e=0.6301
Value
Special type
Size (diameter) Major axes
Rotational properties
Rotation
Precession
Pole position in the space
Density
Ω=124.3991◦
i=0.4466◦ ,
ω=278.6910◦ , M=6.7634◦
S (IV)5
(x=4.60±0.10 km, y=2.29±0.10 km, z=1.92± 0.10 km)5
(x=4.46±0.10 km, y=2.27±0.10 km, z=1.88± 0.10 km)9
5.4 day2, 5
7.4 day2, 5
β = 250 ± 5◦, λ = 63 ± 5◦
λ = 60.6◦8, 9
β = 54.6◦,
α = 258 ± 5◦, δ = 40 ± 5◦
2.5 g cm−3
20
Using the data released by the Minor Planet Center34, 35 and optical data from the ground-
based observational campaign sponsored by the Chinese Academy of Sciences, we determined
Toutatis' orbit with uncertainties on the order of several kilometers. Osculating orbital elements
were calculated for the flyby epoch of 13 December 2012 at 8:30 UTC, where a, e, i, Ω, ω,
and M are the semi-major axis, the eccentricity, the inclination, the longitude of the ascending
node, the argument of perihelion, and the mean anomaly, respectively. β, λ and α, δ are the
longitudes and latitudes of the long axis of the asteroid in the J2000.0 ecliptic and equatorial
coordinate systems, respectively.
21
|
1806.06089 | 1 | 1806 | 2018-06-15T18:34:22 | An absolute sodium abundance for a cloud-free 'hot Saturn' exoplanet | [
"astro-ph.EP"
] | Broad absorption signatures from alkali metals, such as the sodium (Na I) and potassium (K I) resonance doublets, have long been predicted in the optical atmospheric spectra of cloud-free irradiated gas-giant exoplanets1,2,3. However, observations have only revealed the narrow cores of these features rather than the full pressure-broadened profiles4-6. Cloud and haze opacity at the day-night planetary terminator are considered responsible for obscuring the absorption-line wings, which hinders constraints on absolute atmospheric abundances7-9. Here we present an optical transmission spectrum for the 'hot-Saturn' WASP-96b obtained with the Very Large Telescope, which exhibits the complete pressure-broadened profile of the sodium absorption feature. The spectrum is in excellent agreement with cloud-free, solar-abundance models assuming chemical equilibrium. We are able to measure a precise, absolute sodium abundance of log\epsilon_Na=6.9+0.6-0.4, and use it as a proxy to the planet's atmospheric metallicity relative to the solar value (Z_p/Z_\star=2.3+8.9/--1.7). This result is consistent with the mass-metallicity trend observed for solar-system planets and exoplanets10-12. | astro-ph.EP | astro-ph | An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
Nikolay Nikolov1*, David K. Sing1,2, Jonathan J. Fortney3,
Jayesh M. Goyal1, Benjamin Drummond1, Tom M. Evans1,
Neale P. Gibson4, Ernst J. W. De Mooij5,6, Zafar Rustamkulov3,
Hannah R. Wakeford7, Barry Smalley8, Adam J. Burgasser9,
Coel Hellier8, Christiane Helling10,11, Nathan J. Mayne1,
Nikku Madhusudhan12, Tiffany Kataria13, Josef Baines4,
Aarynn L. Carter1, Gilda E. Ballester14, Joanna K. Barstow15,
Jack McCleery4 & Jessica J. Spake1
1 Physics and Astronomy, University of Exeter, Exeter EX4 4QL, UK.
2 Department of Earth and Planetary Sciences, Johns Hopkins University, Baltimore, MD, USA
3 Department of Astronomy and Astrophysics, University of California, Santa Cruz, CA 95064, USA.
4 Astrophysics Research Centre, School of Mathematics and Physics, Queens University Belfast,
Belfast BT7 1NN, UK.
5 School of Physical Sciences, and Centre for Astrophysics & Relativity, Dublin City University,
Glasnevin, Dublin 9, Ireland.
6 Centre for Astrophysics & Relativity, Dublin City University, Glasnevin, Ireland
7 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, Maryland 21218, USA.
8 Astrophysics Group, Keele University, Keele, UK.
9 Department of Physics, University of California, San Diego, CA 92093, USA..
10 Centre for Exoplanet Science, SUPA, School of Physics and Astronomy, University of St. Andrews,
North Haugh, St. Andrews, Fife, KY16 9SS, UK.
11Anton Pannekoek Institute for Astronomy, University of Amsterdam, Science Park 904, 1098 XH
Amsterdam, The Netherlands.
12 Institute of Astronomy, University of Cambridge, Madingley Road, CB3 0HA Cambridge, UK.
13 Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena,
CA, USA.
14 Lunar and Planetary Laboratory, University of Arizona, Tucson, AZ 85721, USA.
15 Physics and Astronomy, University College London, London, UK.
* e-mail: [email protected]
Broad absorption signatures from alkali metals, such as the sodium (Na I)
and potassium (K I) resonance doublets, have long been predicted in the
optical atmospheric spectra of cloud-free irradiated gas-giant exoplanets1,2,3.
However, observations have only revealed the narrow cores of these features
rather than the full pressure-broadened profiles4-6. Cloud and haze opacity
at the day-night planetary terminator are considered responsible for
obscuring the absorption-line wings, which hinders constraints on absolute
atmospheric abundances7-9. Here we present an optical transmission
spectrum for the 'hot-Saturn' WASP-96b obtained with the Very Large
Telescope, which exhibits the complete pressure-broadened profile of the
sodium absorption feature. The spectrum is in excellent agreement with
cloud-free, solar-abundance models assuming chemical equilibrium. We are
able to measure a precise, absolute sodium abundance of 𝐥𝐨𝐠 𝛆𝐍𝐚=𝟔.𝟗,𝟎.𝟒/𝟎.𝟔,
solar value (𝐙𝐩/𝐙⨀=𝟐.𝟑,𝟏.𝟕/𝟖.𝟗). This result is consistent with the mass-
and use it as a proxy to the planet's atmospheric metallicity relative to the
metallicity trend observed for solar-system planets and exoplanets10-12.
WASP-96b
Nikolov et al. 2018
We observed two transits of the hot Saturn WASP-96b (planetary mass
𝑀:= (0.48±0.03) 𝑀B, 𝑅:= (1.20±0.06) 𝑅B, 𝑇HI=1285±40K)13 on 2017
combined cover the wavelength range 3600 to 8200 Å. We used a mask
brightness. Broad slits spanning 22″ along the dispersion and 120″ along the
July 29th and August 22nd UT in photometric conditions, using the 8.2-metre
Unit Telescope 1 of the Very Large Telescope, with the FORS2 spectrograph.
Data were collected in the multi-object-spectroscopy mode using grisms 600B
(blue) and 600RI (red) on the first and second nights, respectively, which
consisting of two broad slits centred on the target and on a reference star of similar
spatial (perpendicular) axis were used in order to minimise slit losses due to
seeing variations and guiding imperfections.
0.120
)
s
R
/
p
0.118
a
clear, Burrows profile, χ2=49
clear, Allard profile, χ2=50
cloud deck, χ2=69
haze, χ2=76
Na
Li
K
10
R
(
o
i
t
a
r
s
u
d
a
r
r
a
i
t
t
s
−
o
−
e
n
a
P
t
l
)
s
R
/
p
R
(
o
i
t
i
a
r
s
u
d
a
r
r
a
t
t
s
−
o
−
e
n
a
P
t
l
b
0.116
0.114
0.112
0.110
0.120
0.120
0.118
0.118
0.116
0.116
0.114
0.114
0.112
0.112
Na
Na
Li
K
Figure 1 Transmission spectrum
of WASP-96b compared to
models. a: Comparison of the
FORS2 observations (black dots
with 1𝜎 vertical error bars; the
horizontal bars indicate spectral bin
widths) with clear3,16, cloudy and
hazy one-dimensional forward
atmospheric models at solar
abundance14 (continuous lines).
The two best-fit models assume a
clear atmosphere with different line
broadening shapes for Na and K
(see text for details). Models with
hazes or clouds (magenta and blue)
predict much smaller and narrower
absorption features. b: Similar to
the upper panel, but showing the
best-fit model obtained from the
retrieval analysis (red line) binned
to the data resolution (red dots),
with the 1s, 2s and 3s confidence
intervals (blue regions).
8
6
4
2
0
i
l
t
h
g
e
H
e
a
c
S
e
r
u
s
s
e
r
P
−2
10
8
6
4
2
0
t
i
l
h
g
e
H
e
a
c
S
e
r
u
s
s
e
r
P
0.110
0.110
.3
.3
H2 Rayleigh
.4
.4
.5
.5
.6
.6
.7
.7
Wavelength (µm)
−2
.8
.8
.9
.9
For each transit, we produced wavelength-integrated "white" and
spectroscopic light curves for WASP-96 and the reference star by integrating the
flux of each spectrum along the dispersion axis. We corrected the light curves for
extinction caused by the Earth's atmosphere by dividing the flux of the target by
the flux of the reference star. We modelled the transit and systematic effects of the
white-light curves by treating the data as a Gaussian Process (GP) and assuming
quadratic limb darkening for the star. The transit parameters: mid-time (𝑇OPQ),
2
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
orbital inclination (𝑖), normalised semi-major axis (𝑎/𝑅*), planet-to-star radius
ratio (𝑅:/𝑅*) and the two limb-darkening coefficients (𝑢U and 𝑢V), were allowed
to vary in the fit to each of the two white-light curves, while the orbital period was
held fixed to the previously determined value. The white-light curves and results
from the modelling are shown in Extended Data Figure 1 and Table 1.
are the elemental abundances with the uncertainties of the host star (dotted lines and
blue regions) and the Sun (dash-dotted lines in grey regions from ref.25).
Figure 2 Retrieved atmospheric properties for WASP-96b. Histograms of
the marginalized posterior distributions from a free retrieval. a, b, Negligible opacity
from clouds and hazes comprises the evidence for a clear atmosphere at the limb
of the planet. c-e, Retrieved elemental abundances in the scale of ref.23, which ranges
from 0 to 12 with the abundance of hydrogen log𝜀[=12. The abundance of Na,
log𝜀\]=6.9,_.'/_.a is the only constrained quantity (c). The vertical continuous and
dotted lines indicate the mean abundances and 1𝜎 uncertainties, respectively. Shown
width of 160 Å. Wavelength-independent systematics were corrected following
standard practices, as detailed in Methods. We allowed only 𝑅:/𝑅* and 𝑢U to
darkening coefficients (𝑢V) were fixed to their theoretical values. To account for
vary. The rest of the system parameters were fixed to their weighted mean values
from the analysis of the two white-light curves, while the quadratic limb-
To obtain the transmission spectrum, we produced 28 and 35
spectroscopic light curves, from the blue and red grisms, respectively, with a
systematics, we marginalised over a grid of polynomials, where the latter
consisted of terms up to second order in air mass and drift of the spectrum across
the detector along both the dispersion and cross-dispersion axes. The resulting
time series are shown in Extended Data Figure 2 and 3.
3
WASP-96b
Nikolov et al. 2018
The measured wavelength-dependent relative planet radii are shown in
Figure 1, which comprises the transmission spectrum of WASP-96b. The
spectrum reveals the absorption signature of the pressure-broadened sodium D
line with wings covering ~6 atmospheric pressure scale heights (1 scale height
corresponds to ~610 km, assuming 𝑇HI=1285K), in a wavelength range from
~5000 to ~7500 Å, and a slope at near-UV wavelengths due to Rayleigh
scattering by molecular hydrogen. The radius measurements around the potassium
feature show no obvious broadened line wing shape or larger absorption at the line
cores.
To interpret the measured transmission spectrum, we first compare it with
clear, cloudy and hazy atmospheric models with solar abundances from ref. 14.
We find that cloud-free models assuming chemical equilibrium best fitted the 49
data points, giving 𝜒V=49 and 𝜒V=50 for a total of 48 degrees of freedom.
Models with clouds and hazes, i.e. 100× enhanced-Rayleigh scattering cross-
section (haze) and 100× enhanced wavelength-independent (cloud) opacity, give
𝜒V of 69 and 76 respectively, and are disfavoured at ~3𝜎 and ~5𝜎 confidence,
respectively (Figure 1). Further details are provided in Methods.
1000.0
100.0
10.0
1.0
0.1
l
)
r
a
o
s
×
(
y
t
i
c
i
l
l
a
t
e
M
⛢♆
WASP−39b
HST+VLT
♃
♄
WASP−96b
VLT
Constraining constituent:
Na
H2O
CH4
0.10
1.00
0.01
Mass (× Jupiter)
10.00
Figure 3 Mass-metallicity diagram for Solar System planets and exoplanets.
Figure 3 Mass-metallicity diagram for solar system planets and exoplanets. Methane
(CH4) and water (H2O) are the two absorbing constituents used to constrain the
Methane (CH4) and water (H2O) are the two absorbing constituents used to constrain
atmospheric metallicity of solar system planets (blue bars) and hot gas-giant exoplanets
the atmospheric metallicity of solar system planets (blue bars) and hot gas-giant
(orange squares with grey error bars), respectively. Absorption lines from atomic Na (red
exoplanets (orange squares with grey error bars), respectively. Absorption lines
triangles) can provide another proxy to exoplanet atmospheric metallicity, which has been
from atomic Na (red triangles) can provide another proxy to exoplanet atmospheric
done by combining three HST and two VLT transits for WASP-39b. With detected and
metallicity, which has been done by combining three Hubble Space Telescope (HST)
resolved pressure-broadened Na line wings, WASP-96b is the first transiting exoplanet for
and two Very Large Telescope (VLT) transits for WASP-39b. With its detected and
which high-precision atmospheric metallicity has been constrained using data only from
resolved pressure-broadened Na line wings, WASP-96b is the first transiting exoplanet
the ground.
for which high-precision atmospheric metallicity has been constrained using data
line indicates a fit to the Solar System gas giants (pale blue symbols indicate
Solar System planets).
only from the ground. Each error bar corresponds to the 1𝜎 uncertainty. The blue
The wing shape of atomic absorption lines is a result of the combined
contribution of the quantum mechanical (natural), thermal (or Doppler) and
collisional (or pressure) broadening mechanisms15. Measurements of the shape of
4
6
7
8
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
pressure-broadened line wings can provide important constraints on the interaction
potentials used in the theory of
stellar and sub-stellar atmospheres16,17. While, such constraints have been obtained
from Na and K absorption lines in the spectra of brown dwarfs18,19, the actual
shape of the profiles for exoplanets remain unconstrained. To assess the detection
of sodium line-broadening we compared the spectrum to models with no
broadened lines. Compared to the best-fit clear-atmosphere model, the narrow-line
model is found to be rejected at the 5.8𝜎 confidence level. This is in contrast to
WASP-39b, WASP-17b and HD209458b, which have previously been classified
as having the clearest atmospheres of the known exoplanets. The latter
transmission spectra are well-explained with narrow alkali features, implying that
the broad absorption wings are masked by clouds and hazes4,5,20,21.
The broad sodium feature measured for WASP-96b therefore provides a
unique opportunity to constrain the pressure-broadened line shape for an
exoplanet atmosphere. We compared the observed spectrum to two cloud-free
models, assuming alkali line-wing shapes from refs. 3 and 16. We find each of
them to be statistically consistent with the data, although the wing profile of ref. 3
is marginally preferred (Figure 1, red and orange models).
covered by three of our measurements. Throughout this letter, we adopt the
or haze opacity, which indicates that the atmosphere of WASP-96b is free of
clouds and hazes at the pressures being probed at the limb. The best-fit
transmission spectrum includes opacity from Na, Li, K and Rayleigh scattering
To further interpret the physical properties of WASP-96b's atmosphere,
we performed a retrieval analysis of the data using the 1D radiative-convective
ATMO model22. We assumed an isothermal atmosphere and allowed the
temperature, radius, opacity from clouds and hazes and elemental abundances of
Na and K to vary. In addition, Li is expected to add opacity at ~6650 Å, which is
astronomical scale of logarithmic abundances of ref. 23, where hydrogen (H) is
defined to be 𝑙𝑜𝑔 𝜀[=12. The abundance of a particular element X is defined as
log𝜀m=log(𝑁m/𝑁[)+12, where 𝑁m and 𝑁[ are the number densities of
elements X and H. Our retrieval analysis finds negligible contributions from cloud
(Figure 1). We obtain a tight constrain of log 𝜖\]= 6.9,_.'/_.a on the sodium
model gives 𝜒V=39 for 42 degrees of freedom.
𝜒V value when excluding the two species is only slightly higher than when they
are included (∆𝜒V=2). However, we include the two species in our retrieval
atmospheric temperature of T= 1710,V__/Us_K, which is somewhat higher,
compared with the planet's equilibrium temperature of THI=1285±40K under
model to marginalise the Na abundance over the possibility of their presence and
estimate upper-limits on their abundances. The abundances of K and Li are also
found to depend on the assumed profile shape of the Na feature. We find an
abundance, which is in agreement with the solar abundance as well as with the
measured sodium abundance in the WASP-96 host star (Figure 2). The best-fit
The current data does not support detections of K or Li, as the minimum
the assumption of zero albedo and uniform day-night heat redistribution 13.
5
WASP-96b
Nikolov et al. 2018
GRIS600B
GRIS600RI
Orbital period (day)
eccentricity
TOPQ (JD)
i (°)
a/R*
R:/R*
uU
uV
A (ppm)
h(cid:135)(cid:136)[(cid:137) (arbitrary normalisation)
hm (arbitrary normalisation)
c_
h(cid:140) (arbitrary normalisation)
cU
TOPQ (JD)
i (°)
a/R*
R:/R*
uU
uV
A (ppm)
h(cid:141)(cid:142)(cid:143) (arbitrary normalisation)
hm (arbitrary normalisation)
c_
h(cid:140) (arbitrary normalisation)
cU
i (°)
a/R*
uU
uV
uU
uV
R:/R*
R:/R*
Weighted mean:
GRIS600RI
3.4252602 (fixed)
0 (fixed)
4013 – 6173 Å
57963.33672,_.___(cid:129)(cid:129)
/_.___(cid:129)'
85.21,_.U(cid:130)/_.U(cid:131)
8.93,_.U(cid:131)/_.U(cid:131)
0.1141,_.__U'
/_.__Us
0.399,_._(cid:129)_
/_._sa
0.148,_._'_
/_._(cid:130)s
501,(cid:130)V/U_(cid:130)
1.01,_.(cid:138)(cid:138)/U.V
−3.45,_.s(cid:130)/_.(cid:130)_
−1.86,U._U/_.(cid:131)(cid:131)
0.9992,_.___Ua
/_.___U(cid:130)
−0.00068,_.___UV
/_.___UV
5268 – 8308 Å
57987.31195,_.___V(cid:138)
/_.___V(cid:138)
85.11,_.UV/_.UV
8.80,_.UV/_.UU
0.1172,_.__U(cid:130)
/_.__U(cid:130)
0.26,_._(cid:131)/_.UU
0.23,_._(cid:138)/_.UV
2301,(cid:130)'U/Ua(cid:138)_
0.50,_.a(cid:138)/_.(cid:131)(cid:130)
2.81,_.(cid:130)a/_.(cid:138)'
8.08,V.s/(cid:129).s
0.9980,_.__U(cid:129)
/_.__Ua
0.00018,_.___V(cid:131)
/_.___V(cid:138)
85.14±0.10
8.84,_.U_/_._(cid:138)
(fixed i, a/R* and TOPQ)
0.1147,_.__U'
/_.__U'
0.435,_._V_
/_._(cid:129)s
0.173,_._VV
/_._'U
(fixed i, a/R* and TOPQ)
0.1168,_.__U'
/_.__Us
0.282,_._(cid:129)V
/_._s(cid:129)
0.250,_._(cid:129)V
/_._s(cid:129)
Extended Data Table 1 System parameters
Parameter
Value
GRIS600B
Heavy element abundance measurements are important to constrain
formation mechanisms of gas-giant exoplanets. Under the core-accretion
paradigm, as the planet mass decreases the atmospheric metallicity increases24,25.
As giant planets accrete H/He-dominated gas as they form, they also accrete
planetesimals26 that enrich their H/He envelopes in metals. A low-mass H/He
envelope has smaller amount of gas for these metals to be mixed into, leading to a
higher metal enrichment compared to the parent star. This is also the scenario for
solar system gas-giants, where metallicity has been constrained from methane
(CH4) abundance from in-situ or infrared-spectroscopy27-30, showing increasing
6
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
enrichment of heavy elements with decreasing mass (Figure 3). Measurements of
H2O abundances have been used to constrain atmospheric metallicities for a small
sample of exoplanets10-12. The measured molecular abundances are used as proxies
to atmospheric metallicities, assuming chemical equilibrium conditions. Using our
measurement of the absolute sodium abundance of WASP-96b, we estimate an
atmospheric metallicity of 𝑍:/𝑍⊙= 2.3,U.(cid:130)/(cid:131).(cid:138), i.e. log (𝑍:/𝑍⊙) =0.4,_.s/_.(cid:130). This is
consistent with the heavy element abundance of the host star 𝑍∗/𝑍⊙ =1.4±0.7
which we estimate using the relation 𝑍∗/𝑍⊙ =10[(cid:135)H/[] , where [Fe/H] =
0.14±0.19. While our WASP-96b measurement is consistent with the solar
system mass-metallicity trend (see Figure 3), we note that additional high-
precision constraints would be necessary to further support or refute a trend for
exoplanets.
WASP-96b is the first exoplanet for which the pressure-broadened wings
of an atomic absorption line (Na I) have been observed, probing deeper layers of
the atmosphere at the limb. This observation has also enabled a precise
atmospheric abundance constraint, using ground-based data alone. Our result
demonstrates that combined with near-UV data, the ~5890Å Na absorption
feature is a valuable probe of exoplanet metallicities accessible to ground-based
telescopes over a wavelength region largely free of contamination by telluric lines.
WASP-96b is the first gas-giant with a detected broad atomic absorption feature
out of approximately 20 exoplanets so far characterized in transmission. This
demonstrates the significant role a future ground-based optical spectrograph,
optimized for transmission spectroscopy, could play. With the clearest atmosphere
of any exoplanet characterized so far, WASP-96b will be an important target for
the upcoming James Webb Space Telescope.
Online Content Any Methods, including any statements of data availability and Nature
Research reporting summaries, along with any additional references and Source Data
files, are available in the online version of the paper at https://doi.org/10.1038/s41586-
018-0101-7.
Received: 27 November 2017; Accepted: 30 January 2018;
Published online: 07 May 2018
1. Seager, S. & Sasselov, D. Theoretical transmission spectra during extrasolar
giant planet transits. Astrophys. J. 537, 916–921 (2000).
2. Sudarsky, D. et al. Albedo and reflection spectra of extrasolar giant planets,
Astrophys. J. 538, 885-903 (2000).
3. Burrows, A. et al. The near-infrared and optical spectra of methane dwarfs and
brown dwarfs. Astrophys. J. 531, 438-446 (2000).
4. Charbonneau, D. et al. Detection of an extrasolar planet atmosphere. Astrophys.
J. 568, 377-384 (2002).
5. Sing, D. K. et al. A continuum from clear to cloudy hot-Jupiter exoplanets
without primordial water depletion. Nature 529, 59-62 (2016).
6. Wyttenbach, A. et al. Hot exoplanet atmospheres resolved with transit
spectroscopy (HEARTS). I. Detection of hot neutral sodium at high altitudes on
WASP-49b, Astron. Astrophys. 602, 36-50 (2017).
7. Fortney, J. J. et al. On the indirect detection of sodium in the atmosphere of the
planetary companion to HD 209458. Astrophys. J. 589, 615-622 (2003).
7
WASP-96b
Nikolov et al. 2018
8. Line, M. R. & Parmentier, V. The influence of nonuniform cloud cover on
transit transmission spectra. Astrophys. J. 820, 78-89 (2016).
9. Benneke, B. & Seager, S. Atmospheric retrieval for super-Earths: uniquely
constraining the atmospheric composition with transmission spectroscopy.
Astrophys. J. 753, 100-122 (2012).
10. Kreidberg, L. et al. A precise water abundance measurement for the hot Jupiter
WASP-43b. Astrophys. J. 793, 27-33 (2014).
11. Line, M. R. et al. No thermal inversion and a solar water abundance for the hot
Jupiter HD 209458b from HST/WFC3 spectroscopy. Astrophys. J. 152, 203-217
(2016).
12. Wakeford, H. et al. HAT-P-26b: A Neptune-mass exoplanet with a well-
constrained heavy element abundance. Science 356, 628-631 (2017).
13. Hellier, C. et al. Transiting hot Jupiters from WASP-South, Euler and
TRAPPIST: WASP-95b to WASP-101b. Mon. Not. R. Astron. Soc. 440, 1982
(2014).
14. Fortney, J. J., Lodders, K., Marley, M. S. & Freedman, R. S. A unified theory
for the atmospheres of the hot and very hot Jupiters: two classes of irradiated
atmospheres. Astrophys. J. 678, 1419–1435 (2008).
15. Collins, G. W. The Fundamentals of Stellar Astrophysics. New York, W.H.
Freeman and Co. 512p. (1989).
16. Allard, N. F. et al. A new model for brown dwarf spectra including accurate
unified line shape theory for the Na I and K I resonance line profiles. Astron.
Astrophys. 411, 473-476 (2003).
17. Burrows, A. & Volobuyev, M. Calculations of the far-wing line profiles of
sodium and potassium in the atmospheres of substellar-mass objects. Astrophys.
J., 583, 985-995 (2003).
18. Johnas, C. M. S. The effects of new Na I D line profiles in cool atmospheres.
Astron. Astrophys. 466, 323-325, (2007).
19. Burgasser, A. J. The spectra of T dwarfs. II. Red optical data. Astrophys. J. 594,
510-524 (2003).
20. Nikolov, N. et al. VLT FORS2 comparative transmission spectroscopy:
detection of Na in the atmosphere of WASP-39b from the ground. Astrophys. J.
832, 191-200 (2016).
21. Helling, Ch. et al. The mineral clouds on HD 209458b and HD 189733b. Mon.
Not. R. Astron. Soc. 460, 855-883 (2016).
22. Tremblin, P. Fingering convection and cloudless models for cool brown dwarf
atmospheres. Astrophys. J. 804, 17-23 (2015).
23. Asplund, M. et al. The Chemical composition of the Sun. Ann. Rev. of Astron.
Astrophys. 47, 481-522 (2009).
24. Fortney, J. J., et al. A Framework for characterizing the atmospheres of low-
mass low-density transiting planets. Astrophys. J. 775, 80-93 (2013).
25. Pollack, J. B. et al. Formation of the giant planets by concurrent accretion of
solids and gas. Icar. 124, 62-85 (1996).
26. Mordasini, C. et al. Extrasolar planet population synthesis. IV. Correlations with
disk metallicity, mass, and lifetime. Astron. Astrophys. 541, 97-120 (2012).
27. Wong, M. H. et al. Updated Galileo probe mass spectrometer measurements of
carbon, oxygen, nitrogen, and sulfur on Jupiter. Icar. 171, 153-170 (2004).
28. Fletcher, L. N. et al. Methane and its isotopologues on Saturn from
Cassini/CIRS observations. Icar. 199, 351-367 (2009).
29. Karkoschka, E. & Tomasko, M. G. The haze and methane distributions on
Neptune from HST-STIS spectroscopy. Icar. 211, 780-797 (2011).
8
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
30. Sromovsky, L. A. et al. Methane on Uranus: The case for a compact CH 4 cloud
layer at low latitudes and a severe CH 4 depletion at high-latitudes based on re-
analysis of Voyager occultation measurements and STIS spectroscopy. Icar.
215, 292-312 (2011).
Acknowledgements. This work is based on observations collected at the European
Organization for Astronomical Research in the Southern Hemisphere under European
Southern Observatory programme 199.C-0467(H). The research leading to these results
received funding from the European Research Council under the European Union's
Seventh Framework Programme (FP7/2007-2013)/ERC grant agreement number 336792.
A.J.B. is a US/UK Fulbright Scholar. J.M.G. and N.J.M. acknowledge support from a
Leverhulme Trust Research Project Grant. J.K.B. is a Royal Astronomical Society
Research Fellow.
Author Contributions N.N. led the design of the VLT FORS2 Large Programme and
the scientific proposal (with contributions from N.P.G., D.K.S. and T.M.E.). N.N. led the
observations, analysis, comparison with forward models and the interpretation of the
result. D.K.S. led the atmospheric retrievals (with contributions from J.M.G.). J.J.F. and
Z.R. provided forward atmospheric models for comparative analysis. B.S. and C.H.
provided elemental abundances of the host star. N.N. wrote the manuscript (with
contributions from D.K.S. and T.M.E.). J.B. and J.McC. performed independent tests on
various parts of the data reduction and analysis as part of their final-year undergraduate
projects under supervision from N.P.G. All authors discussed the results and commented
on the draft.
Author Information Reprints and permissions information is available at
www.nature.com/reprints. The authors declare no competing financial interests. Readers
are welcome to comment on the online version of the paper. Correspondence and requests
for materials should be addressed to [email protected]
9
WASP-96b
Nikolov et al. 2018
Extended Data Table 2 Transmission spectrum
𝒖𝟏
0.484 ± 0.059
0.553 ± 0.059
0.539 ± 0.053
0.493 ± 0.052
0.458 ± 0.054
0.464 ± 0.049
0.479 ± 0.043
0.522 ± 0.039
0.415 ± 0.043
0.422 ± 0.042
0.404 ± 0.038
0.385 ± 0.040
0.337 ± 0.041
0.367 ± 0.044
0.410 ± 0.038
0.288 ± 0.043
0.336 ± 0.040
0.260 ± 0.039
0.349 ± 0.047
0.299 ± 0.042
0.330 ± 0.046
0.303 ± 0.050
0.284 ± 0.044
0.263 ± 0.044
0.270 ± 0.041
0.236 ± 0.049
0.242 ± 0.050
0.169 ± 0.053
0.316 ± 0.044
0.235 ± 0.041
0.194 ± 0.045
0.185 ± 0.046
0.267 ± 0.043
0.250 ± 0.040
0.192 ± 0.038
0.155 ± 0.044
0.087 ± 0.042
0.100 ± 0.052
0.129 ± 0.051
0.104 ± 0.044
0.156 ± 0.048
0.277 ± 0.038
0.102 ± 0.050
0.263 ± 0.046
0.147 ± 0.042
0.190 ± 0.049
0.189 ± 0.049
0.175 ± 0.053
0.214 ± 0.059
𝒖𝟐
0.207
0.219
0.223
0.227
0.231
0.235
0.239
0.242
0.246
0.251
0.255
0.256
0.258
0.262
0.269
0.273
0.275
0.277
0.278
0.283
0.287
0.290
0.291
0.295
0.296
0.296
0.297
0.302
0.313
0.314
0.314
0.314
0.313
0.313
0.313
0.314
0.314
0.314
0.314
0.315
0.315
0.315
0.316
0.316
0.315
0.315
0.315
0.314
0.314
10
𝑹𝒑/𝑹*
0.11479 ± 0.00152
0.11447 ± 0.00182
0.11382 ± 0.00167
0.11554 ± 0.00167
0.11459 ± 0.00205
0.11321 ± 0.00163
0.11505 ± 0.00129
0.11466 ± 0.00123
0.11462 ± 0.00125
0.11361 ± 0.00116
0.11282 ± 0.00100
0.11324 ± 0.00130
0.11440 ± 0.00104
0.11451 ± 0.00138
0.11326 ± 0.00112
0.11318 ± 0.00101
0.11317 ± 0.00096
0.11444 ± 0.00089
0.11365 ± 0.00095
0.11361 ± 0.00089
0.11322 ± 0.00102
0.11541 ± 0.00095
0.11499 ± 0.00095
0.11653 ± 0.00084
0.11685 ± 0.00092
0.11648 ± 0.00086
0.11607 ± 0.00086
0.11545 ± 0.00081
0.11487 ± 0.00096
0.11648 ± 0.00099
0.11457 ± 0.00096
0.11499 ± 0.00091
0.11553 ± 0.00100
0.11284 ± 0.00089
0.11532 ± 0.00081
0.11442 ± 0.00117
0.11189 ± 0.00078
0.11311 ± 0.00090
0.11483 ± 0.00096
0.11428 ± 0.00085
0.11463 ± 0.00097
0.11572 ± 0.00094
0.11357 ± 0.00081
0.11405 ± 0.00096
0.11473 ± 0.00075
0.11312 ± 0.00109
0.11341 ± 0.00096
0.11336 ± 0.00123
0.11287 ± 0.00117
𝝀 (Å)
3500−4013
4013−4093
4093−4173
4173−4253
4253−4333
4333−4413
4413−4493
4493−4573
4573−4653
4653−4733
4733−4813
4813−4893
4893−4973
4973−5053
5053−5133
5133−5213
5213−5293
5293−5373
5373−5453
5453−5533
5533−5613
5613−5693
5693−5773
5773−5853
5853−5933
5933−6013
6013−6093
6093−6173
6173−6253
6253−6333
6333−6413
6413−6493
6493−6573
6573−6653
6653−6733
6733−6813
6813−6973
6973−7053
7053−7133
7133−7213
7213−7293
7293−7373
7373−7453
7453−7533
7533−7693
7693−7773
7773−7853
7853−7933
7933−8013
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
METHODS
Observations.
We observed two transits of WASP-96b with the FOcal Reducer and
Spectrograph (FORS2)31 attached on the Unit Telescope 1 (UT1, Antu) of the
Very Large Telescope (VLT) at the European Southern Observatory (ESO) on
Cerro Paranal in Chile as part of Large Program 199.C-0467 (PI: Nikolov). We
used similar observing setup and strategy to our VLT FORS2 Comparative
Transmission Spectroscopy of WASP-39b and WASP-31b22,32.
During the two transits, we monitored the flux of WASP-96 and one
detector (MIT), which is a mosaic of two chips. We positioned the instrument
field of view such that each detector imaged one source. The field of view was
monitored without guiding interruptions during the full observing campaigns. To
reference star at photometric conditions. The reference star, known as 2MASS
00041885-4716309 is the only bright source in the FORS2 field of view and is
located at an angular separation of 5'.3 away from the target. Fortunately, the
reference is of similar colour and brightness, which reduced the effect of
differential colour extinction. For example, the magnitude differences (target
minus reference) from the PPMXL33 catalogue are ΔB=−0.46, ∆R=−0.49,
∆I=−0.5. We observed both transits with the same slit mask and the red
improve the duty cycle, we made use of the fastest available read-out mode (200
kHz, ~30s). During both nights, we ensured that the Longitudinal Atmospheric
(hereafter blue and 600B), which covers the spectral range from 3600 to 6200 Å
at a resolving power of R ~ 600. The field of view rose from an airmass 1.43 to
1.08 and set to an airmass of 1.16. The seeing oscillated around 0.5" during the
first 3.5 hours and gradually increased to 1.2" at the end of the observation. We
GRIS600RI (hereafter red and 600RI), which covers the range from 5400 to
8200 Å, in combination with the GG435 filter to isolate the first order. The field
of view rose from an air mass of 1.23 to 1.08 and set at an airmass of 1.36. The
seeing varied between 0.3" and 0.5" as measured from the cross-dispersion
collected a total of 89 exposures for ~5h with an integration times adjusted
between 120 and 230s.
Dispersion Corrector (LADC) is in its neutral position, i.e. inactive.
During the first night, we used the dispersive element GRIS600B
During the second night, we exploited the dispersive element
profiles of the spectra. We monitored WASP-96 and the reference star for
~5h20m and collected a total of 233 spectra with integration times between 30
and 80s.
Calibrations and data reduction.
We performed data reduction and analysis using a customized IDL pipeline22.
We started by subtracting a bias frame and by applying a flat field correction to
the raw images. We computed a master bias and flat field by obtaining the
median of 100 individual frames. Cosmic rays were identified and corrected
following the routine detailed in ref. 34. We extracted 1D spectra using IRAF's
APALL task. To trace the stars, we used a fit of a Chebyshev polynomial of two
parameters. We performed background correction by subtracting the median
11
WASP-96b
Nikolov et al. 2018
located away from the spectral trace. We found that aperture radii of 14 and 12
pixels and sky regions 21:72 and 23:74 minimize the dispersion of the out-of-
background from the stellar spectrum for each wavelength, computed from a box
transit flux of the band-integrated white light curves for the blue and red
observations, respectively.
4000
4500
5000
5500
Wavelength (Å)
6000
5500
6000
6500
7000
7500
8000
100
10
1
1.04
1.02
1.00
0.98
1.00
WASP−96
Reference star
2017/07/29
GRIS600B
Raw light curves
A: WASP−96/Reference
best−fit GP model
(transit+systematics)
WASP−96
Reference star
2017/08/22
GRIS600RI
3
0
1
×
s
n
o
r
t
c
e
e
o
t
o
h
P
l
t
n
a
t
s
n
o
c
+
x
u
l
f
w
a
r
d
e
z
i
l
a
m
r
o
N
t
n
a
t
s
n
o
c
+
x
u
l
f
e
v
i
t
l
a
e
r
d
e
s
i
l
a
m
r
o
N
)
m
p
p
(
l
i
a
u
d
s
e
R
100
10
1
1.01
1.00
0.99
0.98
1.00
0.99
0.98
0.97
500
0
−500
detrended light curve
0.99
B: best−fit transit model
0.98
0.97
500
0
−500
A/B: common−mode
correction
−3
−2
−1
0
1
Time since mid−transit (hr)
Time since mid−transit (hr)
−2
2
−3
−1
0
1
2
Extended Data Figure 1 VLT FORS2 stellar spectra and white-light curves. Left and right column panels
show the GRIS600B (blue) and GRIS600RI (red) data sets, respectively. First row: example stellar spectra
used for relative spectrophotometric calibration. The dashed lines indicate the wavelength region used to
produce the white-light curves. Second row: normalised raw light curves of both sources. Third row:
Normalised relative target-to-reference raw flux along with the marginalized GP model (A), detrended transit
light curve and model (B), and the common-mode correction (A/B), Fourth row: Best-fit light curve residuals
and 1𝜎 error bars, obtained by subtracting the marginalized transit and systematics models from the relative
target-to-reference raw flux. The two light curve residuals show dispersion of 78 and 201 parts-per-million
(ppm), respectively.
12
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
We performed a wavelength calibration of the extracted stellar spectra
using spectra of an emission lamp, obtained after each of the two transit
observations with a mask identical to the science mask, but with slit widths of 1".
We established a wavelength solution for each of the two stars with a low-order
Chebyshev polynomial fit to the centres of a dozen lines, which we identified by
performing a Gaussian fit. To account for displacements during the course of
each observation and relative to the reference star, we placed the extracted
spectra on a common Doppler-corrected rest frame through cross-correlation. All
spectra were found to drift in the dispersion direction to no more than 2.5 pixels,
with instrument gravity flexure being the most likely reason.
Example spectra of WASP-96 and the reference star are shown in
Extended Figure 1. We achieved typical signal-to-noise ratios (S/N) of 315 and
280 per pixel for the central wavelength of the blue grism and 313 and 257 for
We produced white light curves from 4013 to 6173 and from 5293 to 8333 Å
the red grism, respectively. We then used the extracted spectra to produce band-
integrated white and spectroscopic light curves for each source and transit by
summing up the flux along the dispersion axis in each bandpass.
White light curve analysis.
systematics) in the data:
for the blue and red observations, respectively. We corrected the raw flux of the
target by dividing by the raw flux of the reference star. This correction removes
the contribution of Earth's atmospheric transparency variations, as demonstrated
in Extended Data Figure 1. We modelled the white light transits and instrumental
systematics simultaneously by treating the data as a Gaussian process (GP)35-37.
We performed the GP analysis using the Python GP library George38-41. Under
the GP assumption, the data likelihood is a multivariate normal distribution with
a mean function 𝜇 describing the deterministic transit signal and a covariance
matrix Κ that accounts for stochastic correlations (i.e. poorly-constrained
𝑝( 𝑓 𝜃,𝛾) = 𝒩 (𝜇,Κ),
where 𝑝 is the probability density function, 𝑓 is a vector containing the flux
measurements, 𝜃 is a vector containing the mean function parameters, 𝛾 is a
function containing the covariance parameters, 𝒩 is a multivariate normal
distribution. We defined the mean function 𝜇 as follows:
𝜇(𝑡,𝑡;𝑐_,𝑐U,𝜃)= 𝑐_+𝑐U𝑡Τ(𝑡;𝜃),
where 𝑡 is a vector of all central exposure time stamps in Julian Date (JD), 𝑡 is a
and divided by the standard deviation, 𝑐_ and 𝑐U describe a linear baseline trend,
Τ(θ) is an analytical expression describing the transit and 𝜃=(𝑖,𝑎/
𝑅*,𝑇‹›fi,𝑅fl/𝑅*,𝑢U,𝑢V), where 𝑖 is the orbital inclination, 𝑎/𝑅*, is the
normalized semi-major axis, 𝑇‹›fi is the central transit time, 𝑅fl/𝑅* is the planet-
to-star radius ratio, 𝑢U and 𝑢V are the linear and quadratic limb darkening
coefficients. To obtain an analytical transit model Τ, we used the formulae found
vector containing all standardized times, i.e. with subtracted mean exposure time
in ref. 42. We fixed the orbital period to its value from ref. 13 and fitted for the
remaining system parameters.
13
WASP-96b
Nikolov et al. 2018
Extended Data Figure 2 Spectrophotometric light curves from grism 600B offset by a constant amount for
clarity. First panel: raw target-to-reference flux. Second panel: common-mode corrected light curves and the
transit and systematics models, with the highest evidence. Third panel: detrended light curves and the transit
model with the highest evidence. Fourth panel: residuals with the 1-s error bars. The dashed lines indicate the
median residual level, with dotted lines indicating the dispersion and the reached percentage of the theoretical
photon noise limit (blue).
14
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
parameter (𝑢U,𝑢V) quadratic law and computed the values of the coefficients
We accounted for the stellar limb-darkening by adopting the two-
using a three-dimensional stellar atmosphere model grid43. In these calculations,
we adopted the closest match to the effective temperature, surface gravity and
metallicity of the exoplanet host star found in ref. 13. The choice of a quadratic
versus a more complex law, e.g. the four-parameter non-linear law of ref. 44 was
motivated by the study of refs. 45 and 46, where the two-parameter law has been
demonstrated to introduce negligible bias on the measured properties of transiting
systems similar to WASP-96. In addition, the quadratic law requires much
shorter computational time for the relevant transit light curve. We computed the
theoretical limb-darkening by fitting the limb-darkened intensities of the 3D
stellar atmosphere models, factored by the throughputs of the blue and red
grisms.
Κ=𝜎›V𝛿›–+𝑘›–,
where 𝜎› are the photon noise uncertainties, 𝛿›– is the Kronecker delta function
and 𝑘›– is a covariance function. We assumed the white noise term was the same
for all data points and allowed it to vary as a free parameter, 𝜎‡. For the
covariance function, we choose to use the Matérn 𝜈 = 3/2 kernel with the
spectral dispersion and cross-dispersion drifts 𝑥 and 𝑦, respectively as input
and the speed of the rotation angle 𝑧 (see Extended Data Figure 4). As with the
variables, FWHM, measured from the cross-dispersion profiles of the 2D spectra
The covariance matrix is defined as
linear time term, we also standardised the input parameters prior to the light
curve fitting. We chose to use the dispersion and cross-dispersion drifts for both
observations, and combined them with the FWHM for the blue data and the
speed of the rotation angle for the red data, respectively. Our choice was justified
based on the fact that those combinations of input parameters gave well-behaved
residuals. The covariance function then was defined as:
𝑘›–=ΑV(1+ 3𝐷›–) exp ( − 3𝐷›–),
where Α is the characteristic correlation amplitude and
+(𝑧›−𝑧–)V
+(𝑦›−𝑦–)V
𝐷›–= (𝑥›−𝑥–)V
𝜏(cid:190)V
𝜏‰V
𝜏…V
where 𝜏…, 𝜏‰ and 𝜏(cid:190) are the correlation length scales and the hatted variables are
standardised. We allowed parameters 𝑋= (𝑐_,𝑐U,𝑇‹›fi,𝑖,𝑎/𝑅*,𝑅fl/𝑅*,𝑢U,𝑢V)
and 𝑌= (Α,𝜏…,𝜏‰,𝜏(cid:190)) to vary and fixed the orbital period 𝑃 to its literature
value13. We adopted uniform priors for 𝑋 and log-uniform priors for 𝑌.
To marginalise the posterior distribution 𝑝 (𝜃,𝛾 𝑓) ∝
𝑝 (𝑓 𝜃,𝛾) 𝑝 (𝜃,𝛾) we made use of the Markov-Chain Monte Carlo (MCMC)
software package emcee40. We identified the maximum likelihood solution using
the Levenberg–Marquardt least-squares algorithm47 and initialised three groups
of 150 walkers close to that maximum. We run groups one and two for 350
samples and the third group had 4500 samples. Before running for the second
group we re-sampled the positions of the walkers in a narrow space around the
15
WASP-96b
Nikolov et al. 2018
We computed the weighted mean values of the orbital inclination and
semi-major axis and repeated the fits. In the second fit we allowed only the
position of the best walker from the first run. This extra re-sampling step was
useful because otherwise some of the walkers can start in a low likelihood area of
parameter space and would require more computational time to converge. Transit
models for each of the two observations computed using the marginalized
posterior distributions are shown in Extended Data Figure 1 and the relevant
parameter values are reported in Extended Data Table 1. We find residual
dispersion of 79 and 203 parts-per-million for the blue and red light curves,
respectively. Both values are found to be within 80% of the theoretical the
photon noise limit.
to the weighted mean values and the central times were fixed to the values
determined from the first fit.
Spectroscopic light curve analysis.
We produced spectroscopic light curves by summing the flux of the target and
planet-to-star radius ratio (𝑅fl/𝑅*) and the two limb darkening coefficients (𝑢U
and 𝑢V) to vary, while the orbital inclination and semi-major axis were held fixed
reference star in bands with a width of 160Å. The sodium D lines at 5890 and
5896Å fall inside the spectral range of each grism, within their overlapping
region from 5300 to 6200Å. We centred the set of bins for each night on the
pairs of bins, covering the O2 A and B bands from 7594 to 7621 and from 6867
to 6884 Å, respectively to increase the signal to noise ratio of the corresponding
sodium line, which gave the advantage of obtaining two radius measurements
identical in wavelength coverage within that overlapping region. We merged two
light curves. The very first band in the blue grism was also enlarged with the
same motivation. With these customizations at hand we produced a total of 63
light curves.
Common mode factors
The FORS2 spectroscopic light curves are known to exhibit wavelength
independent (common mode) systematic, as demonstrated from our Comparative
Transmission Spectroscopy studies22,32 and other FORS2 results48-50. This makes
the instrument outstanding for transmission spectroscopy, with enormous
potential to explore the diversity of exoplanet atmospheres. We established the
wavelength independent systematic using the band-integrated white light curves
for each of the two nights. We simply divided the white-light transit light curves
by a transit model. We computed the transit model using the weighted mean
values of the orbital inclination and normalized semi-major axis from both nights
and assumed the central times found from the white light analysis. The values for
the relative radius and the limb darkening coefficients were identified by
repeating the GP fit where the transit central time, orbital inclination and semi-
major axis were held fixed to the weighted-mean values. The fitted relative radii
and limb-darkening coefficients are reported at the end of Extended Data Table
1. The common mode factors for each night are shown in Extended Data Figure 1
along with a schematic explanation of the full white-light curve analysis.
16
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
Extended Data Figure 3 Same as Extended Data Figure 2 but for grism 600RI.
17
WASP-96b
Nikolov et al. 2018
Spectroscopic light curve fits
We modelled the spectroscopic light curves using a two-component function that
takes into account the systematics and transit simultaneously. The transit model
was computed using the analytical formulae of ref. 42, as for the white light
curve, but we allowed in the fits only the relative planet radius (𝑅fl/𝑅∗) and the
linear limb darkening coefficient (𝑢U) to vary. Fitting only for the linear limb-
uncertainty of 𝑢V is large and consistent with the theoretical prediction. We
significantly change we chose to fix 𝑢V and to fit only for 𝑢U. Prior to fitting the
darkening coefficient is a standard practice for ground-based observations that
has been proven to generally perform well51-56. Similar to our WASP-39b study
with FORS222, we also fit for both limb-darkening coefficients and found that the
interpret this as an indication for insufficient constraining power of the data for
the non-linear coefficient. However, as the transmission spectra did not
spectroscopic light curves, we removed the common mode factors from each
night by dividing each of the spectroscopic light curves to the corresponding
common-mode light curve of the same night. We accounted for the systematics
using a low-order polynomial (up to a second degree with no cross terms) of
dispersion and cross-dispersion drift, air mass, FWHM variations and the rate of
change of the rotator angle (Extended Data Figure 4). We produced all possible
combinations of detrending variables and performed separate fits with each
combination with the systematics function included in the two-component model.
For each attempted function, we computed the Akaike Information Criterion
(AIC)57 to estimate statistical weight of the model depending on the number of
degrees of freedom. We marginalized the resulting relative radii and linear limb-
darkening coefficient following ref. 58. We chose to rely on the AIC instead of
other information criteria, e.g. the Bayesian Information Criterion (BIC)59,
because the AIC selects more complex models, resulting in more conservative
error estimates. We note that a marginalization over multiple systematics
functions relies on the assumption of equal prior weights for each tested model.
This assumption is valid for simple polynomial expansions of basis parameters,
as the ones in our study. We found systematics models, parameterized with a
linear air mass, dispersion drift and FWHM terms to result in the highest
evidence.
Prior to each fit, we set the uncertainties of each spectrophotometric
channel to values that are based on the expected photon noise with additional
component from readout noise. We determined best-fit models using a
Levenberg-Marquardt least-squares algorithm and rescaled the uncertainties of
the fitted parameters with the dispersion of the residuals. All residual outliers
larger than 3𝜎 were excluded from the analysis. We found that for each
spectroscopic light curve ≤3 data points were removed. We also assessed the
modelling the binned variance with 𝜎V= (𝜎‡)V/𝑁 +(𝜎)V relation, where
𝜎‡ is the uncorrelated white noise component, N is the number of measurements
in the bin and 𝜎 is the red noise component. We find white and red noise
dispersion in the range from 400 to 1000 and from 20 to 80 ppm, respectively.
levels of correlated residual red noise, following the methodology of ref. 60 by
18
12
s
s
a
m
r
i
A
1.1
1.2
1.3
1.4
0.8
0.6
0.4
0.2
0.0
−0.2
0.6
0.4
0.2
0.0
−0.2
−0.4
)
x
p
(
t
n
e
m
e
c
a
l
p
s
i
d
n
o
i
s
r
e
p
s
i
d
−
s
s
o
r
C
)
x
p
(
t
n
e
m
e
c
a
l
p
s
i
d
n
o
i
s
r
e
p
s
i
D
9
8
7
6
5
4
)
x
p
(
M
H
W
F
)
n
i
m
/
g
e
d
(
d
e
e
p
S
e
l
g
n
A
n
o
i
t
a
t
o
R
0.5
0.4
0.3
0.2
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
GRIS600B
−2
−1
1
Time since mid−transit (hr)
0
−2
−1
1
Time since mid−transit (hr)
0
−2
−1
1
Time since mid−transit (hr)
0
−2
−1
1
Time since mid−transit (hr)
0
a)
c)
e)
g)
i)
2
2
2
2
GRIS600RI
−2
−2
−2
−2
−1
Time since mid−transit (hr)
0
1
−1
Time since mid−transit (hr)
0
1
−1
Time since mid−transit (hr)
0
1
−1
Time since mid−transit (hr)
0
1
b)
d)
f)
h)
j)
2
2
2
2
1.1
1.2
1.3
1.4
0.2
0.1
0.0
−0.1
−0.2
s
s
a
m
r
i
A
)
x
p
(
t
n
e
m
e
c
a
l
p
s
i
d
n
o
i
s
r
e
p
s
i
d
−
s
s
o
r
C
)
x
p
(
t
n
e
m
e
c
a
l
p
s
i
d
n
o
i
s
r
e
p
s
i
D
0.10
0.05
0.00
−0.05
−0.10
4.5
4.0
3.5
3.0
2.5
2.0
1.5
0.5
0.4
0.3
)
x
p
(
M
H
W
F
)
n
i
m
/
g
e
d
(
d
e
e
p
S
e
l
g
n
A
n
o
i
t
a
t
o
R
−2
−1
1
Time since mid−transit (hr)
0
2
Extended Data Figure 4 Light curve auxiliary variables. Shown are air mass (a, b), drifts
along the cross-dispersion (c, d) and dispersion axes (e, f), full width at half maximum (g,
h) and the rate of change of the rotation angle (i, j) of the VLT FORS2 observations. Left
and right columns refer to the GRIS600B and GRIS600RI observations, respectively.
Extended Data Figure 4 Light curve auxiliary variables. Shown are air mass (a, b), drifts along the cross-
dispersion (c, d) and dispersion axes (e, f), full width at half maximum (g, h) and the rate of change of the
rotation angle (i, j) of the VLT FORS2 observations. Left and right columns refer to the GRIS600B and
GRIS600RI observations, respectively.
−2
−1
Time since mid−transit (hr)
1
0
2
13
19
WASP-96b
Nikolov et al. 2018
The measured wavelength dependent relative radii and corresponding
light curves are plotted in Figure 1 and Extended Data Figures 2, 3, 4, 5, Table 2
and comprise the transmission spectrum of WASP-96b. We used the overlapping
wavelength region to combine the two observations together and computed the
weighted mean of both datasets. We detect a marginally significant 1.4𝜎
7.2±5.2 × 10,'. This level of variation is consistent with the photometric
of 𝜎<9.2 × 10,' (ref. 13).
variability, associated with the active regions on the star surface with variability
difference in the transit depths of the light curves from the two observations of
clear, Burrows profile, χ2=65
clear, Allard profile, χ2=66
cloud deck, χ2=85
haze, χ2=92
Na
Li
K
4
2
0
t
i
l
h
g
e
H
e
a
c
S
e
r
u
s
s
e
r
P
−2
−4
600B
600RI
−6
0.3
0.4
0.5
0.6
0.7
0.8
0.9
Wavelength (µm)
measurements from grism 600B (blue dots) and 600RI (red dots) along with the 1𝜎 uncertainties,
Extended Data Figure 5 Transmission spectrum of WASP-96b with indicated relative
Extended Data Figure 5 Transmission spectrum of WASP-96b. Indicated are the relative radius
radius measurements from grism 600B (blue dots) and 600RI (red dots) compared to the
same set of models as in Figure 1.
compared to the same set of models as in Figure 1.
Modelling of the atmosphere of WASP-96b.
14
)
s
R
/
p
R
(
o
i
t
i
a
r
s
u
d
a
r
r
a
t
t
s
−
o
−
e
n
a
P
t
l
0.120
0.118
0.116
0.114
0.112
0.110
15
16
17
18
19
20
21
22
23
24
First we compare the observed transmission spectrum to the models
based on ref. 16, which included a self-consistent treatment of radiative transfer
and chemical equilibrium of neutral and ionic species. Chemical mixing ratios
and opacities were computed assuming solar metallicity and local chemical
equilibrium, accounting for condensation and thermal ionization but not
photoionization61-63. Transmission spectra were also calculated using 1D T-P
profiles for the dayside, as well as an overall cooler planetary-averaged profile.
In addition to the atmospheric models of a clear atmosphere, a simplified
treatment of the scattering and absorption have been incorporated in those
models to simulate the effect of small particle haze aerosols and large particle
cloud condensates at optical and near-infrared wavelengths. In the case of haze,
20
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
cloud deck we included a wavelength-independent cross-section, which was a
Rayleigh scattering opacity (𝜎 = 𝜎_(𝜆/𝜆_),') has been assumed with a cross-
section which was 1,000× the cross-section of molecular hydrogen gas (𝜎_ =
5.31 × 10,V(cid:130)cmV at 𝜆_ = 3500Å; ref. 61). To include the effects of a flat
factor of 100× the cross-section of molecular hydrogen gas at 𝜆_ = 3500Å;
with a single parameter responsible for their vertical offset5,22,34. The 𝜒Vand BIC
freedom for each model determined by 𝜈=𝑁−𝑚, where 𝑁 is the number of
measurements and 𝑚 the number of free parameters in the fit.
statistic quantities were computed for each model with the number of degrees of
models within the wavelength bins and fitted these theoretical values to the data
Similar to our previous studies, we obtained the average values of the
(see Fig. 1).
y
t
i
s
n
e
D
m
a
r
g
o
i
t
s
H
d
e
z
i
l
a
m
r
o
N
1.00
0.80
0.60
0.40
0.20
0.00
0
2
4
6
8
10
log(Na/K)
respectively. The solar value is indicated with the blue continuous line.
Extended Data Figure 6 Na/K ratio Histogram of the marginalized
posterior distribution of the Na to K ratio for WASP-96b. Shown are
Extended Data Figure 6 Na/K ratio Histogram of the marginalized posterior distribution
of the Na to K ratio for WASP-96b. Shown are the median and 1 sigma levels with orange
continuous and dotted lines, respectively. The solar value is indicated with the blue
continuous line.
the median and 1𝜎 levels with orange continuous and dotted lines,
line from 4500 to 6800 Å with cloud free atmosphere models, assuming two
26. We find no statistically significant difference between the two with 𝜒V=22
and 𝜒V=30 for 35 d.o.f. Further observations at higher signal to noise and
individual shapes of the pressure broadened profile. The first profile had the
shape detailed in ref. 3 and the second line shape from the formalism of ref. 18,
We also compared the observed transmission spectrum around the Na D
resolution will be necessary to distinguish between the two line-wing shapes.
Retrieval analysis
We performed a retrieval8,9,64 analysis using the 1D radiative-convective
equilibrium ATMO model24,65-68. The code includes isotropic multi-gas Rayleigh
scattering, H2-H2 and H2-He collision-induced absorption, as well as opacities for
all major chemical species taken from the most up-to-date high-temperature
21
25
26
27
28
29
30
31
32
33
34
WASP-96b
Nikolov et al. 2018
sources, including: H2, He, H2O, CO2, CO, CH4, Na, K, Li, Rb and Cs 66,69. Hazes
and clouds are respectively treated as parameterized enhanced Rayleigh
scattering and grey opacity70, which is similar to the approach of ref. 16 and is
detailed in ref. 69. ATMO uses the correlated-k approximation with the random
overlap method to compute the total gaseous mixture opacity, which has been
shown to agree well with a full line-by-line treatment66.
We first performed a free retrieval analysis, allowing a fit to the
abundances of Na, K and Li, cloud and haze opacity, the planet's radius and
atmospheric temperature. Na, K and Li were included as these elements are
expected to add significant opacity in the wavelength region of our observations.
We assumed that these gasses are well-mixed vertically in the atmosphere.
1000.0
100.0
10.0
1.0
0.1
r
a
t
s
Z
/
t
e
n
a
p
Z
l
⛢♆
WASP−39b
HST+VLT
♄
WASP−96b
VLT
♃
Constraining constituent:
Na
H2O
CH4
0.01
0.10
1.00
10.00
Mass (× Jupiter)
For all free parameters in our model we adopted uniform priors with the
error bar represents the 1𝜎 uncertainty. The blue line indicates a fit to the Solar System gas giants (pale blue
following ranges: 250−3000K for the temperature, 1−1.5𝑅B for the radius,
10,U_ to 10(cid:129) for the cloud and 10,U_ to 10(cid:129) for the haze opacity. For the
between −2 and 12. Our retrieval analysis proceeded by first identifying the
minimum 𝜒V solution using nonlinear least-squares optimization and then
the Gelman-Rubin statistic for each free parameter was within 1% of unity,
chains had found a 𝜒V below the median 𝜒V value of the chain. Finally, we
Extended Data Figure 7 Heavy element enrichment of exoplanets relative to their stars as a function of
Extended Data Figure 7 Heavy element enrichment of exoplanets relative to their stars
mass. Plotted are the solar system planets (blue bars) and gas-giant exoplanets (grey and orange symbols). Each
as a function of mass. The blue line indicates a fit to the solar system gas giants.
symbols indicate Solar System planets).
mixing ratios of chemical species other than H and He we adopted uniform priors
marginalizing over the posterior distribution using differential-evolution Markov
chain Monte Carlo71. A total of 12 chains for 30,000 steps each, were ran until
showing that the chains were well-mixed and had reached a steady state. We
discarded a burn-in phase from all chains corresponding to the step at which all
22
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
combined the remaining samples into a single chain, forming our posterior
distributions. We summarize the results from the free retrieval in Figure 2.
Elemental abundances of Na and Li were obtained for the host star from
high-resolution optical spectroscopy with typical S/N of ∼100:1, as detailed in
ref. 13. Because the potassium resonance doublet is located outside of the red
limit of the high-resolution spectra, it was not possible to obtain abundance
measurement.
the limb with precise sodium abundance consistent with the solar value at ~1𝜎.
measurement at ~0.73𝜇m shows a slightly larger radius, though it is only ~1.8𝜎
the solar value at the ~2𝜎 confidence (Figure 1 and 2). This most likely is a
above our best-fit model. Our retrieval analysis shows a marginalized posterior
distribution bounded only on the upper bound with median value consistent with
The best-fitting retrieved spectrum suggests a cloud-free atmosphere at
The spectrum shows no definitive evidence of the potassium feature. A
consequence of the limited wavelength range, covering only ~2/3 of the
potassium feature, as well as the diluting effect of the O2-A band, which partially
covers this wavelength regime. The limited constraining power of the K feature
allows for a three-case scenario for the abundance of K in the atmosphere of
WASP-96b. In both cases the atmosphere of the planet is clear at the limb with
sodium at solar abundance. In the first scenario, the abundance of potassium is
sub-solar, leading to a missing potassium feature in the transmission spectrum. In
chemical equilibrium, Na and K can form condensates (e.g Na2S and KCl) at
temperatures lower that ~1300 and 1000 K, respectively. However, given that
~1300 to 1700 K this is unlikely.
the retrieved temperature in the region probed by our observation lies between
In the second scenario, the potassium can be ionized by the UV radiation
of the host star, which would also lead to missing potassium line cores. The
pressure-broadened line wings would still be present, because they would
originate from deeper layers, where the UV radiation would be able to penetrate.
We estimate the relative sodium-to-potassium ratio from our free retrieval
analysis, finding highly uncertain value, which prevents definitive conclusion
(Extended Data Figure 6). Future observations of multiple transits could help
resolve the two-case scenario.
A third scenario would require stratified atmosphere consisting of a low-
altitude potassium layer followed by a cloud cover on top, above which is located
a clear atmosphere consisting sodium at solar abundance. The sodium layer
would need to be deep enough to allow for pressure-broadened line wings in the
transmission spectrum.
We note a large degeneracy between the aerosol clouds/hazes and the Na
abundance, as higher cloud opacity levels can be fit with increased Na
abundances. This degeneracy can be seen in the posterior distribution of the
retrieval (see Extended Data Figure 8), and the marginalized posterior
distributions are shown in Figure 2 (i.e. the distribution shown in Extended Data
Figure 8 integrated along each axis). The degeneracy is accounted for in our
retrieval modelling by marginalizing the Na abundance over the other fit model
23
WASP-96b
Nikolov et al. 2018
parameters, which includes the effects of clouds and hazes. An aerosol-free
model where the near-UV transmission spectra is dominated by H2 Rayleigh
scattering helps determine the lower limit to the Na abundance. The upper limit
for the Na abundance is sensitive to the line profile shape, as very high aerosol
opacity levels probe significantly lower pressure levels, which affects the line
profile. For example, for a clear atmosphere the transmission spectra at 4000 Å
probes pressures of 20 mbar and is dominated by molecular hydrogen at that
wavelength. In a model where the haze opacity is 1000x stronger than molecular
hydrogen (the pink haze model shown in Figure 1), the pressure probed is
~1000x lower or ~0.02 mbar. These lower pressures affect the calculation of the
pressure-broadened sodium line profile. As described in ref. 3, using semi-
classical impact theory the half-width of the Lorentzian sodium line core is
determined by the effective collision frequency, which itself is a product of the
H2- perturber density (among other factors). We find that in the described
scenario with very high sodium abundances and high cloud/haze opacities, very
low pressures (such as 0.02 mbar) are probed and the sodium line profile
becomes too narrow to be able to fit the wide profile as seen in the data, even at
arbitrarily high abundances.
Extended Data Figure 8 Posterior
distribution of the retrieved cloud
opacity versus the sodium abundance.
VMR, vertical mixing ratio. The median
and 1𝜎 measured parameters are
indicated with continuous lines and the
red dot marks the intercept for clarity.
The colour scale shows the normalized
density of the samples in the MCMS run.
(cid:2)(cid:1)(cid:1)(cid:1),
(cid:6)(cid:1)(cid:1)
(cid:5)(cid:1)(cid:1)
(cid:4)(cid:1)(cid:1)
(cid:3)(cid:1)(cid:1)
(cid:1)
(cid:4)(cid:5)(cid:7)(cid:8)(cid:6)(cid:9)(cid:10)
We acknowledge that the 1D model does not account for horizontal
advection which may affect the composition. This can break the chemical
equilibrium conditions, leading to horizontally-constant non-equilibrium
abundances of various species e.g., CH4 (ref. 72), which could potentially affect
the observed transmission spectrum. However, an analysis of such effects is
beyond the scope of the present letter.
The presence of sodium and potassium in the atmospheric spectra of
irradiated gas giant exoplanets has proved puzzling, with some of the spectra
showing both or either of the two features without a clear trend with the planetary
properties5. Primordial abundance variation along with atmospheric processes,
such as condensation, photochemistry and photoionisation have been
hypothesized to be responsible for the observed alkali variation. Elemental
abundance constraints with ground-based instruments such as FORS2 could help
24
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
identify the role of some of those processes from statistically large samples of
exoplanet spectra.
relation of atmospheric metallicity with abundance of detected absorption
features. Currently, such estimations have been obtained using the 1.4 µm water
Exoplanet atmospheric metallicities have been estimated using scaling
absorption feature. Recently, ref. 73 have obtained such measurement from a
combined HST + VLT spectroscopy. We performed a retrieval analysis assuming
chemical equilibrium in the atmosphere and determine the metallicity of WASP-
96b's atmosphere, finding consistency with the metallicity of the host star and
the mass-metallicity trend established for solar system giant planets (Figure 3).
We acknowledge that such estimations can largely be inaccurate due to the
scaling metallicity assumption with elemental abundance and round-off the
precision of the WASP-96b's metallicity to one significant figure. The current
census of exoplanet metallicity measurements exhibits a large scatter across the
full mass-metallicity diagram, which hampers definitive conclusion regarding a
trend. Similar to ref. 74 we plot the relative planet to star heavy element
enrichment as a function of the planet mass instead of the planet metallicity alone
(Extended Data Figure 7). We convert 𝑍:/𝑍⨀ to 𝑍:/𝑍∗ assuming a scaling
approximation with the parent star iron abundance of the form 𝑍:/𝑍∗=10[(cid:135)H/[],
where 𝑍⨀=0.014 and propagate the uncertainty of [Fe/H]. It remains a future
work to improve the statistics on the mass versus metallicity diagram with
additional precise measurements. We therefore, consider it pre-mature to claim
hypothetical trends or emerging patterns and relevant interpretation.
Code availability. Publicly available custom codes were used for the Gaussian process
modelling george (http://dfm.io/george/current/user/gp/) and emcee code
(http://github.com/ dfm/emcee). The MCMC retrieval analyses were performed using the
publicly available package exofast (http://astroutils.astronomy.ohio-state.edu/exofast).
The ATMO code used to compute the atmosphere models is currently proprietary. We
have opted not to make the customized IDL codes used to produce the spectra publicly
available owing to their undocumented intricacies.
Data availability. The data is within standard proprietary period of one year and will
become publicly available on ESO archive (July and August) in 2018. Reduced data
products and models used in this study are available Supplementary Information.
31. Appenzeller, I. et al. Successful commissioning of FORS1 - the first optical
instrument on the VLT. Msngr. 94, 1-6 (1998).
32. Gibson, N. P. et al. VLT/FORS2 comparative transmission spectroscopy II:
Confirmation of a cloud deck and Rayleigh scattering in WASP-31b, but no
potassium? MNRAS, 467, 4591-4605 (2017).
33. Roeser, S. et al. The PPMXL catalogue of positions and proper motions on the
ICRS. combining USNO-B1.0 and the two micron all sky survey (2MASS). AJ,
139, 2440-2447 (2010).
34. Nikolov, N. et al. Hubble Space Telescope hot Jupiter transmission spectral
survey: a detection of Na and strong optical absorption in HAT-P-1b. MNRAS,
437, 46-66 (2014).
25
WASP-96b
Nikolov et al. 2018
35. Gibson, N. P. et al. A Gaussian process framework for modelling instrumental
systematics: application to transmission spectroscopy. MNRAS, 419, 2683-
2694 (2012).
36. Evans, T. M. et al. An ultrahot gas-giant exoplanet with a stratosphere. Natur.
548, 58-61 (2017).
37. Nikolov N. et al. Hubble PanCET: An isothermal day-side atmosphere for the
bloated gas-giant HAT-P-32Ab. MNRAS 474, 1705-1717 (2018).
38. Ambikasaran, S. et al. Fast direct methods for Gaussian processes. ITPAM, 38,
252-265 (2015).
39. Foreman-Mackey, D. George: Gaussian process regression, Astrophysics Source
Code Library. ascl.soft11015 (2015).
40. Foreman-Mackey, D. et al. emcee: The MCMC hammer. PASP, 125, 306-312
(2013).
41. Foreman-Mackey, D., corner.py: Scatterplot matrices in Python. Journal of
Open Source Software, 1, 24 (2016).
42. Mandel, K. & Agol, E. Analytic lightcurves for planetary transit searches,
Astrophys. J. 580, L171-175 (2002).
43. Magic, Z. et al. The Stagger-grid: A grid of 3D stellar atmosphere models. IV.
Limb darkening coefficients. Astron. & Astrophys., 573, 90-100 (2015).
44. Claret, A. A new non-linear limb-darkening law for LTE stellar atmosphere
models II K at several surface gravities. Geneva and Walraven systems:
Calculations for -5.0<=log[M/H]<=+1 2000K<=Teff<=50,000 K at several
surface gravities. Astron. & Astrophys. 401, 657-660 (2003).
45. Espinoza, N. & Jordan, A. Limb darkening and exoplanets: testing stellar model
atmospheres and identifying biases in transit parameters. MNRAS, 450, 1879-
1899 (2015).
46. Espinoza, N. & Jordan, A. Limb darkening and exoplanets - II. Choosing the
best law for optimal retrieval of transit parameters. MNRAS, 457, 3573-3581
(2016).
47. Markwardt, C. B. Non-linear least-squares fitting in IDL with MPFIT. ASPC,
411, 251-254 (2009).
48. Sedaghati, E. et al. Detection of titanium oxide in the atmosphere of a hot
Jupiter. Natur. 549, 238-241 (2017).
49. Sedaghati, E. et al. Potassium detection in the clear atmosphere of a hot-Jupiter.
FORS2 transmission spectroscopy of WASP-17b. Astron. Astrophys., 596, 47-
61 (2016).
50. Lendl, M. FORS2 observes a multi-epoch transmission spectrum of the hot
Saturn-mass exoplanet WASP-49b. Astron. Astrophys. 587, 67-81 (2016).
51. Southworth, J. Homogeneous studies of transiting extrasolar planets - I. Light-
curve analyses. MNRAS, 386, 1644-1666 (2008).
52. Mallonn, M. et al. Transmission spectroscopy of the inflated exo-Saturn HAT-P-
19b. Astron Astrophys, 580, 60-71 (2015).
53. Stevenson, K. B. A Search for water in the atmosphere of HAT-P-26b using
LDSS-3C. Astrophys. J., 817, 141-151 (2016).
54. Nikolov, N. et al. HST hot-Jupiter transmission spectral survey: haze in the
atmosphere of WASP-6b. MNRAS, 447, 463-478 (2015).
55. Rackham, B. et al. ACCESS I: An optical transmission spectrum of GJ 1214b
reveals a heterogeneous stellar photosphere. Astrophys. J., 834, 151-172 (2017).
56. Huitson, C. M. Gemini/GMOS transmission spectral survey: complete optical
transmission spectrum of the hot Jupiter WASP-4b. Astron J., 154, 95-113
(2017).
26
An absolute sodium abundance for a cloud-free
'hot Saturn' exoplanet
57. Akaike, H. A new look at the statistical model identification. ITAC, 19, 716-723
(1974).
58. Gibson, N. P. Reliable inference of exoplanet light-curve parameters using
deterministic and stochastic systematics models. MNRAS, 445, 3401-3414
(2014).
59. Schwarz, G. Estimating the dimension of a model. AnSta, 6, 461-464 (1978).
60. Pont, F. et al. The effect of red noise on planetary transit detection. MNRAS,
373, 231-242 (2006).
61. Lodders, K., Alkali element chemistry in cool dwarf atmospheres. Astrophys J.,
519, 793-801 (1999).
62. Lodders, K. & Fegley, B. Atmospheric chemistry in giant planets, brown
dwarfs, and low-mass dwarf stars. i. carbon, nitrogen, and oxygen. Icar. 155,
393-424 (2002).
63. Freedman, R. S. et al. Line and mean opacities for ultracool dwarfs and
extrasolar planets. Astrophys. J. Suppl. Ser. 174, 504-513 (2008).
64. Madhusudhan, N. & Seager, S. A temperature and abundance retrieval method
for exoplanet atmospheres. Astrophys. J. 707, 24-39 (2009).
65. Amundsen, D. S. et al. Accuracy tests of radiation schemes used in hot Jupiter
global circulation models. Astron. Astrophys., 564, 59-75 (2014).
66. Amundsen, D. S. et al. Treatment of overlapping gaseous absorption with the
correlated-k method in hot Jupiter and brown dwarf atmosphere models. Astron.
Astrophys., 598, 97-107 (2017).
67. Tremblin, P. et al. Advection of potential temperature in the atmosphere of
irradiated exoplanets: a robust mechanism to explain radius inflation. Astrophys.
J., 841, 30-38 (2017).
68. Drummond, B. The effects of consistent chemical kinetics calculations on the
pressure-temperature profiles and emission spectra of hot Jupiters. Astron.
Astrophys., 594, 69-84 (2016).
69. Goyal, J. M. A library of ATMO forward model transmission spectra for hot
Jupiter exoplanets. MNRAS, 474, 5158-5185 (2017).
70. Lecavelier des Etangs, A. et al. Rayleigh scattering in the transit spectrum of HD
189733b. Astron. Astrophys. 481, L83–L86 (2008).
71. Eastman, J., Gaudi, B. S. & Agol, E. EXOFAST: a fast-exoplanetary fitting suite
in IDL. Publ. Astron. Soc. Pacif. 125, 83–112 (2013).
72. Drummond, B. et al. Observable signatures of wind-driven chemistry with a
fully consistent three-dimensional radiative hydrodynamics model of HD
209458b. Astrophys. J. Let. (in rev.).
73. Wakeford, H. et al. The Complete transmission spectrum of WASP-39b with a
precise water constraint. Astron. J., 155, 29-43 (2018).
74. Thorngren, Daniel P. et al. The mass-metallicity relation for giant planets.
Astrop. J., 831, 64-78, (2016).
27
|
1005.1633 | 2 | 1005 | 2010-05-22T03:56:53 | Observations of Mass Loss from the Transiting Exoplanet HD 209458b | [
"astro-ph.EP",
"astro-ph.SR"
] | Using the new Cosmic Origins Spectrograph (COS) on the {\it Hubble Space Telescope (HST)}, we obtained moderate-resolution, high signal/noise ultraviolet spectra of HD 209458 and its exoplanet HD 209458b during transit, both orbital quadratures, and secondary eclipse. We compare transit spectra with spectra obtained at non-transit phases to identify spectral features due to the exoplanet's expanding atmosphere. We find that the mean flux decreased by $7.8\pm 1.3$% for the C II 1334.5323\AA\ and 1335.6854\AA\ lines and by $8.2\pm 1.4$% for the Si III 1206.500\AA\ line during transit compared to non-transit times in the velocity interval --50 to +50 km s$^{-1}$. Comparison of the C II and Si III line depths and transit/non-transit line ratios shows deeper absorption features near --10 and +15 km s$^{-1}$ and less certain features near --40 and +30--70 km s$^{-1}$, but future observations are needed to verify this first detection of velocity structure in the expanding atmosphere of an exoplanet. Our results for the C II lines and the non-detection of Si IV 1394.76\AA\ absorption are in agreement with \citet{Vidal-Madjar2004}, but we find absorption during transit in the Si III line contrary to the earlier result. The $8\pm 1$% obscuration of the star during transit is far larger than the 1.5% obscuration by the exoplanet's disk. Absorption during transit at velocities between --50 and +50 km s$^{-1}$ in the C II and Si III lines requires high-velocity ion absorbers, but models that assume that the absorbers are high-temperature thermal ions are inconsistent with the COS spectra. Assuming hydrodynamic model values for the gas temperature and outflow velocity at the limb of the outflow as seen in the C II lines, we find mass-loss rates in the range (8--40)$\times 10^{10}$ g s$^{-1}$. | astro-ph.EP | astro-ph |
Draft November 8, 2018
OBSERVATIONS OF MASS LOSS FROM THE TRANSITING
EXOPLANET HD 209458b 1
Jeffrey L. Linsky
JILA, University of Colorado and NIST, 440 UCB Boulder, CO 80309-0440
[email protected]
Hao Yang
JILA, University of Colorado and NIST, 440 UCB Boulder, CO 80309-0440
[email protected]
Kevin France
CASA, University of Colorado, 593 UCB Boulder, CO 80309-0593
[email protected]
Cynthia S. Froning
CASA, University of Colorado, 593 UCB Boulder, CO 80309-0593
[email protected]
James C. Green
CASA, University of Colorado, 593 UCB Boulder, CO 80309-0593
[email protected]
John T. Stocke
CASA, University of Colorado, 593 UCB Boulder, CO 80309-0593
[email protected]
Steven N. Osterman
CASA, University of Colorado, 593 UCB Boulder, CO 80309-0593
-- 2 --
[email protected]
ABSTRACT
Using the new Cosmic Origins Spectrograph (COS) on the Hubble Space Telescope
(HST), we obtained moderate-resolution, high signal/noise ultraviolet spectra of
HD 209458 and its exoplanet HD 209458b during transit, both orbital quadra-
tures, and secondary eclipse. We compare transit spectra with spectra obtained
at non-transit phases to identify spectral features due to the exoplanet's expand-
ing atmosphere. We find that the mean flux decreased by 7.8± 1.3% for the C II
1334.5323 A and 1335.6854 A lines and by 8.2 ± 1.4% for the Si III 1206.500 A
line during transit compared to non-transit times in the velocity interval -- 50 to
+50 km s−1. Comparison of the C II and Si III line depths and transit/non-
transit line ratios shows deeper absorption features near -- 10 and +15 km s−1
and less certain features near -- 40 and +30 -- 70 km s−1, but future observations
are needed to verify this first detection of velocity structure in the expanding
atmosphere of an exoplanet. Our results for the C II lines and the non-detection
of Si IV 1394.76 A absorption are in agreement with Vidal-Madjar et al. (2004),
but we find absorption during transit in the Si III line contrary to the earlier
result. The 8 ± 1% obscuration of the star during transit is far larger than the
1.5% obscuration by the exoplanet's disk. Absorption during transit at velocities
between -- 50 and +50 km s−1 in the C II and Si III lines requires high-velocity
ion absorbers. Assuming hydrodynamic model values for the gas temperature
and outflow velocity at the limb of the outflow as seen in the C II lines, we find
mass-loss rates in the range (8 -- 40)×1010 g s−1. These rates assume that the
carbon abundance is solar, which is not the case for the giant planets in the solar
system. Our mass-loss rate estimate is consistent with theoretical hydrodynamic
models that include metals in the outflowing gas.
Subject headings: planets and satellites: atmospheres -- planets and satellites:
individual (HD 209458b) -- planets and satellites: physical evolution -- stars:
individual (HD 209458) -- ultraviolet: stars
1Based on observations made with the NASA/ESA Hubble Space Telescope, obtained from the Data
Archive at the Space Telescope Science Institute. STScI is operated by the Association of Universities for
Research in Astronomy, Inc., under NASA contract NAS 5-26555. These observations are associated with
the HST GTO program #11534.
-- 3 --
1.
INTRODUCTION
Mazeh et al. (2000) monitored the radial velocities of the G0 V star HD 209458 to de-
rive the orbital parameters and minimum mass of its transiting planet HD 209458b. Very
shortly thereafter, accurate measurements of the mass and radius of HD 209458b were ex-
tracted from radial velocity and photometric transit observations by Henry et al. (2000) and
Charbonneau et al. (2000). Table 1 lists the presently accepted properties of HD 209458 and
HD 209458b cited by Knutson et al. (2007). HD 209458b is likely the best-studied Jupiter-
like exoplanet located very close to its host star. As a result of its proximity, HD 209458b
receives very strong incident radiation (see estimate in Table 1) and stellar wind flux from
its host star.
HD 209458b is the first transiting planet for which atmospheric absorption was observed
in the resonance lines of Na I (Charbonneau et al. 2002) and subsequently in lines of H I, C II,
O I, and others. Using the G140M grating of Space Telescope Imaging Spectrograph (STIS)
on the Hubble Space Telescope HST with its resolution of ∼ 30 km s−1, Vidal-Madjar et al.
(2003) found that the Lyman-α line flux was reduced by 15 ± 4% during transit at velocities
between -- 130 and +100 km s−1. Based on this first detection of H I absorption at or
above the planet's Roche lobe, they concluded that hydrogen is escaping from the planet.
Vidal-Madjar et al. (2004) then used the low-resolution G140L grating of STIS to detect an
absorption depth at midtransit of 5 ± 2% for the unresolved Lyman-α line, 13 ± 4.5% for
the O I 1304 A multiplet, and 7.5 ± 3.5% for the C II 1335 A doublet. Their detection of
absorption in the O I and C II lines at or above the Roche lobe confirmed significant mass
loss from the planet by hydrodynamic outflow. They argued that reduction in the Lyman-
α line flux at ≈ 100 km s−1 from line center indicates high velocities of neutral hydrogen
atoms. Additionally, more absorption on the blue side of the Lyman-α emission line than
the red side indicates outflow towards the observer and away from the planet. A subsequent
observation of HD 209458 using the HST Advanced Camera for Surveys (ACS) showed a
reduction in the unresolved Lyman-α line flux of (8.0 ± 5.7)% during transits, consistent
with the STIS G140L and G140M results, thereby confirming mass loss from the planet's
large exosphere (Ehrenreich et al. 2008).
There have been a number of reanalyses of the data and futher discussions of the STIS
transit G140M observations by Ben-Jaffel (2007), Ben-Jaffel (2008), Vidal-Madjar et al.
(2008) and others. There is a consensus that HD 209458b has an extended exosphere and
is losing mass, but important details are not yet understood. Such questions as: whether
the exosphere size is larger or smaller than the projected Roche lobe, what are the outflow
speed and mass-loss rate, and whether or not the mass loss from the planet produces a
cometlike tail are not yet answered with the existing STIS and ACS observations. While
-- 4 --
the STIS G140M Lyman-α spectrum obtained by Vidal-Madjar et al. (2003) provides some
velocity information, higher spectral resolution and signal/noise (S/N) are needed to obtain
the velocity structure and optical depth of the outflowing gas in Lyman-α and other spectral
lines.
The large abundance of hydrogen and large flux in the Lyman-α line make this line a
prime candidate for studying exoplanet outflows, but there are uncertainties (many of them
unique) in measuring the exoplanet's mass-loss rate only from this line. These include the
modest S/N of the existing STIS G140M grating data, possible variability of the stellar
Lyman-α profile at the time of transit that cannot be directly measured, broad interstellar
absorption in the core of the line that prevents measurements of low velocity absorption
during transit, and the time-varying geocoronal and interplanetary emission in the line core
that cannot be completely corrected for.
Absorption of Lyman-α photons during transit over the velocity range -- 130 to +100
km s−1 raises the question of the origin of the high-velocity hydrogen that is difficult to ex-
plain by a thermal plasma at moderate temperature. Holmstrom et al. (2008) proposed that
the high-velocity neutral hydrogen can be produced by charge exchange between stellar wind
protons and neutral hydrogen in the planet's outflow. Such energetic neutral atoms (ENAs)
are seen in the solar system and likely occur where the stellar and planetary winds interact
between HD 209458b and its host star. If this were the only possible explanation, then the
Lyman-α profile during transit provides useful information on the stellar wind but ambiguous
information on mass loss from the planet (Holmstrom et al. 2008; Murray-Clay et al. 2009).
However, Lecavelier des Etangs et al. (2008b) argued that the observed Lyman-α profile and
the high-velocity hydrogen atoms can be simply explained by stellar radiation pressure and
the ENA explanation required a peculiar stellar wind model.
The Lyman-α transit observations by Vidal-Madjar et al. (2003, 2004) stimulated a
number of interpretive papers concerning the mechanisms for mass loss and lifetime of
a Jupiter-like planet close to its host star [e.g., Vidal-Madjar et al. (2003); Baraffe et al.
(2004); Koskinen et al. (2007); Holmstrom et al. (2008); Koskinen et al. (2010)] and hy-
drodynamic models [e.g., Lammer et al. (2003); Lecavelier des Etangs et al. (2004); Yelle
(2004); Jaritz et al. (2005); Tian et al. (2005); Garcia Munoz (2007); Schneiter et al. (2007);
Murray-Clay et al. (2009)]. As discussed below, the theoretical models differ in their treat-
ment of the amount of available radiative energy input (photoionization and otherwise) to
drive the hyrdodynamic blow-off, the location of the heating, the presence of metals in the
outflow, and the relative amount of cooling by expansion, thermal conduction, and radiation.
These models predict very different mass-loss rates. Which of these models are consistent
with the spectrally resolved metal lines observed during transit?
-- 5 --
With the objectives of measuring an accurate mass-loss rate from the planet and testing
different physical models for the mass loss, one needs far better data, which requires a new
scientific instrument. One needs higher S/N spectra with higher spectral resolution and
sufficient sensitivity to study many spectral lines formed in the extended atmosphere and
wind of the exoplanet. In particular, it is important to use spectral lines other than Lyman-α
to avoid unique difficulties in the analysis of this line. As we describe below, the new Cosmic
Origins Spectrograph (COS) on HST is well designed for this task.
2. OBSERVATIONS AND DATA REDUCTION
Installed on the HST during Servicing Mission 4 in 2009 May, COS is a high-throughput
ultraviolet (UV) spectrograph optimized for point sources. Descriptions of the on-orbit
performance characteristics of COS will be presented by Green et al. (2010, in preparation)
and Osterman et al.
(2010, in preparation). During GTO Program 11534, we observed
HD 209458 with both the G130M and G160M gratings of the COS far-UV channel to obtain
moderate-resolution spectra covering the 1140 -- 1790 A spectral region.
In this paper, we
analyze the spectra obtained with the G130M grating, which includes the spectral range
1140 -- 1450 A. The G160M observations are presented by France et al. (2010). The COS line-
spread function (LSF) in the far-UV is not a simple Gaussian, but it can be approximated
by a Gaussian with a resolution of 17,000-18,000 (16.7-17.6 km s−1) with extended wings
(Ghavamian et al. 2010).
We have tested for the effects of the extended wings by comparing the C II 1335.6854 A
emission line of α Cen A (G2 V) observed with the STIS E140H grating (3 km s−1 resolution)
convolved with a Gaussian (17 km s−1 resolution) and with the COS LSF. In the line wings
at 50 -- 70 km s−1, the Gaussian broadens the profile by less than 1 km s−1, but the COS LSF
broadens the profile by 2 -- 6 km s−1. Since both the transit and non-transit line profiles are
broadened in the same way, their ratio is not changed by the broad wings of the COS LSF.
The broadening of the intrinsic spectral lines by the COS LSF (compared to the Gaussian
profile) is less than 1/3 of the spectral resolution element and thus not significant.
Table 2 summarizes the individual G130M exposures obtained at transit (near orbital
phase 0.00), first quadrature (phase 0.25), secondary eclipse (phase 0.50), and second quadra-
ture (phase 0.75). The total exposure times for these phases are 4946.8 s, 7941.7 s, 4956.7
s, and 7796.0 s, respectively. The orbital phases listed in Table 2 are computed using the
ephemeris of Knutson et al. (2007). Transit occurs between orbital phases 0.982 and 0.018
(Wittenmyer et al. 2005) as indicated by the optical light curve. Since the duration of transit
is about 3.22 hours, HST could observe HD 209458 during the same transit on two sequential
-- 6 --
spacecraft orbits.
The S/N in the COS spectra can be lower than predicted by photon statistics as a result
of two instrumental effects. The hex pattern formed at the intersection of the microchannel
plate pores imposes a ±4% pattern on the spectral signal, and shadows of the wire grid
on the detector decrease the signal as much as 20% at certain locations on the detector
faceplate. We checked the locations of the known wire grid shadows and found that the
spectra of the C II, Si III, and Si IV lines studied in this paper are not affected by the
wire grids. Both effects, which are fixed at the detector, were mitigated by moving the
grating and the resultant spectrum to different positions on the detector. In this program,
we observed at four different grating offset positions identified by the beginning and ending
wavelengths of the gap between the two detector faceplate elements listed in the last column
of Table 2. In principle, division of a transit spectrum by an non-transit spectrum at the
same grating position should cancel the signal errors introduced by both instrumental effects
to show the true spectral difference produced by the planet in front of the star. In practice,
the cancellation will not be perfect because of possible time variations in the stellar spectrum
and grating mechanisms that do not return to the same position every time. The grating
mechanism positioning errors are not fully understood at this time. We found that the
wavelength solutions for spectra observed at different grating offset positions are not identical
and could be off by 1 -- 2 pixels. We registered each observed spectrum with others at the
same grating setting by cross-correlating the interstellar feature in the C II 1334.5323 A
line. These registrations typically involved displacements by only 1 -- 2 pixels, far less than
a resolution element with no significant degradation in the spectral resolution. We then
co-added the summed spectra for the four grating offset positions using the same cross-
correlation technique. Also, within 500 pixels of the end of each detector segment (about
5 A for G130M), there are additional instrumental artifacts in the spectrum. For this reason,
we did not analyze portions of the spectrum near the ends of a detector faceplate. The
data were processed with the COS calibration pipeline, CALCOS2 v2.11b (2009-09-08) and
combined with a custom IDL coaddition procedure.
Figure 1 shows a portion of the G130M spectra obtained during transit and all non-
transit phases (secondary eclipse and both quadratures). The spectra are very similar, but
there are subtle differences that provide information on the exoplanet's atmosphere and
mass-loss rate. In this paper, we compare transit and non-transit spectra of lines of C II,
Si III, and Si IV. The H I Lyman-α line cannot be analyzed because geocoronal emission
through the large aperture of COS completely dominates any stellar componant. The O I
2We
refer
the
reader
to
the
cycle
18 COS Instrument Handbook
for more details:
http://www.stsci.edu/hst/cos/documents/handbooks/current/cos cover.html.
-- 7 --
lines at 1302, 1304, and 1306 A were also not analyzed because time-varying geocoronal
emission in these lines renders comparison of transit and non-transit spectra uncertain and
because the lines were located close to the end of a detector faceplate.
3. RESULTS
3.1.
Si IV Line
A critical question is whether the emission line fluxes of the host star were significantly
different at the times when the transit and non-transit spectra were obtained, since late-type
stars like the Sun often show time-variable emission in lines formed in their outer atmospheres
(e.g., Rottman 2006). Figure 2 shows the co-added transit and non-transit spectra for the
Si IV resonance line (1393.76 A). Since the Si IV line is formed in highly ionized gas which is
not likely to be in the atmosphere or extended wind of the exoplanet, the very similar Si IV
fluxes during times of transit and non-transit indictate that the average flux in the stellar
emission lines during our observations was nearly constant. The difference spectrum panel
in Figure 2 confirms that the transit and non-transit spectra of the Si IV line are identical
to within the noise level.
To make the comparison of the transit and non-transit Si IV line profiles more quantita-
tive, we compute the mean transit/non-transit flux ratio over the velocity range -- 50 to +50
km s−1. Beyond this velocity range, the fluxes are less than 15% of the peak flux, and the
S/N in each spectral resolution element is small. For the 100 km s−1 wide velocity interval
centered on 0 km s−1, the mean flux ratio is 0.998. To estimate the error in the mean flux
ratio, we have computed flux ratios for many velocity intervals contained within the -- 50
to +50 km s−1 velocity range that are between 60 and 95 km s−1 wide. The dispersion of
these flux ratios indicates that the standard deviation in the mean flux ratio for the -- 50 to
+50 km s−1 velocity range is ±0.014. We therefore conclude that the mean stellar fluxes
in the other emission lines were the same during transit and non-transit times, and do not
rescale the transit and non-transit fluxes. At positive velocities, there are regions where
the transit/non-transit ratio may differ from unity at or above one standard deviation. We
discuss this in Section 3.4.
3.2. C II Lines
The C II lines at 1334.5323 A and 1335.6854 A are bright emission lines formed in the
chromospheres of solar-type stars like HD 209458. On October 2, 2009, we observed the C II
-- 8 --
lines during transit with the G130M grating at four grating positions specified by the gap
wavelengths listed in Table 2. For each of these grating positions, there are corresponding
observations at both quadratures and a secondary eclipse obtained at different times.
Figure 3 shows the co-added spectra of the C II 1334.5323 A resonance line and the
C II 1335.6854 A line. The spectra were smoothed with a 5-pixel boxcar, which is somewhat
smaller than the on-orbit G130M spectral resolution element of 7 -- 8 pixels. The transit and
non-transit spectra are plotted separately and ratioed. The deep minimum observed in the
C II 1334.5323 A line at -- 6.60 km s−1 (heliocentric) is interstellar absorption due to the Eri
cloud located within 3.5 pc of the Sun (Redfield & Linsky 2008). Since the radial velocity
of HD 209458 is -- 14.8 km s−1, the -- 6.6 km s−1 interstellar absorption is centered to the red
of the centroid of the stellar emission line.
Both C II lines show less flux during transit than non-transit times, although there are
small velocity intervals where the transit/non-transit ratio exceeds unity. To test whether
the ratios in these velocity intervals are real or noise, we have co-added the profiles of the two
C II lines (including the portion of the 1334 A line with interstellar absorption) to get transit
and non-transit profiles with twice the signal and higher S/N. The co-added C II profiles
are included in Figure 3, together with difference and ratio profiles for each C II line and
the co-added line. The co-added C II line shows absorption during transit at all velocities
between -- 50 and +50 km s−1, except at zero velocity. The deep absorption features for the
co-added C II line are interesting and will be discussed in Section 3.4.
Using the same method that we used for the Si IV line, we find that the mean flux ratio
in the -- 50 to +50 km s−1 velocity interval is 0.924 ± 0.022 for the C II 1334.53 A line and
0.921 ± 0.015 for the C II 1335.69 A line. For the co-added C II line, the mean flux ratio is
0.922 ± 0.013.
3.3.
Si III Line
Figure 2 shows the co-added transit and non-transit spectra for the Si III resonance line
at 1206.500 A. The difference and ratio spectra show absorption at all velocities between -60
and +60 km s−1. There are also deep absorption features which we will discuss in Section
3.4.
Using the same method that we used in analyzing the Si IV and C II lines, we find
that the mean flux ratio in the -- 50 to +50 km s−1 band is 0.918 ± 0.014, the same within
errors as previously seen for the C II lines. Our result for the mean absorption depth during
transit of 8.2 ± 1.4% is very different from the non-detection absorption depth of 0.0+2.2
−0.0%
-- 9 --
obtained by Vidal-Madjar et al. (2004) from the STIS G140L spectrum and confirmed by
Ben-Jaffel & Sona Hosseini (2010). We suspect that the difference between our detection of
Si III absorption during transit and the previous non-detection is due to time variability.
Charge transfer between protons and Si+ ions leads to a substantially higher abundance of
Si++ than is the case for collisional ionization equilibrium. Baliunas & Butler (1980) found
that Si++ is the dominant ionization stage in coronal plasmas at temperatures above 20,000 K
and is important at 15,000 K. Thus, we are not surprized to see a substantial amount of
Si++ at the limb of the exoplanet's outflow where protons from the solar wind and planetary
escape are present to ionize Si+, but the amount of Si++ could vary appreciably with changes
in the stellar wind, planetary mass-loss rate, and temperature of the gas escaping from the
planet's exosphere.
3.4.
Is There Velocity Structure in the Planet's Mass Loss?
Until now we have discussed the mean transit/non-transit flux ratios for the entire line
cores between -- 50 and +50 km s−1, but there are indications of departures from these mean
ratios which indicate that the absorption in the planet's mass loss depends on velocity. To
test this idea, we compare in Figure 4 the difference and ratio profiles for the Si III and
co-added C II lines. We find deep absorption features near -- 10 and +15 km s−1 in both
lines that are larger than the indicated errors shown in Figures 2 and 3. Since the C II and
Si III lines are likely formed in the same region of the exosphere, we include in Figure 4
the co-addition of the two profiles with typical error bars for comparison with theoretical
models. The increase in S/N with the co-addition of the C II and Si III spectra allows us to
make more definitive statements about the velocity structure. The dips near -- 10 and +15
km s−1 are more than twice as large as the errors per spectral resolution element (about 17
km s−1) and are most likely real. In addition, the ratios near -- 40 and +30 -- 70 km s−1 are
also low, but the line fluxes decrease rapidly away from line center and these low ratios are
less certain. Future observations with higher S/N are needed to confirm this first detection
of velocity structure in the expanding atmosphere of an exoplanet. We also note that there
are absorption features near +20 and +40 km s−1 in the transit/non-transit ratio for the
Si IV line which are greater than the noise level. Future observations are also needed to test
the reality of these features.
-- 10 --
3.5. Does HD 209458b have a Comet-like Tail?
Schneider et al. (1998) proposed that giant exoplanets located close to their host star
will have comet-like tails of ionized gas produced by the interaction of the stellar wind and
ions excaping from the planet's exosphere, and that these tails could produce absorption
when observed in front of the star. Vidal-Madjar et al. (2003) then predicted that stellar
Lyman-α radiation pressure on hydrogen atoms escaping from the planet's exosphere would
form a comet-like tail trailing the planet. The presence of gas in an extended tail should
produce more absorption at and beyond egress than during ingress phases. Schneiter et al.
(2007) calculated possible cometary wakes and the time evolution of Lyman-α absorption
from ingress to beyond egress. We searched our C II and Si III spectra for line profile
differences between the ingress and egress phases to see whether these ions could be present
in significant amounts in a tail, but we found no differences larger than the noise level. Future
observations with higher S/N are required to verify the predictions of comet-like tails.
4. MASS-LOSS RATE IN A THERMAL PLASMA OUTFLOW
The ≈ 8 ± 1% obscuration of the star in the C II and Si III lines during transit far
exceeds the 1.5% obscuration by the exoplanet's disk, indicating absorption by an extended
atmosphere or mass loss that is optically thick in the C II and Si III lines. Our result
for the C II lines is consistent with the 7.5+3.6
−3.4% absorption depth that Vidal-Madjar et al.
(2004) obtained for the unresolved C II lines, but very different from the absorption depth
of 0.0+2.2
−0.0% for the
Si IV line is consistent with our result.
−0.0% that they obtained for the Si III line. Their absorption depth of 0.0+6.5
Since the size of the obscuring material (optically thick in these lines) is similar to the size
of the exoplanet's Roche lobe (Ben-Jaffel & Sona Hosseini 2010), one possible explanation
for the transit data is that the Roche lobe is filled by mass loss from the planet. Another
possible explanation is that the exoplanet's escaping gas forms an extended cometary tail
blown out by the stellar wind, as suggested by hydrodynamical simulations (Schneiter et al.
2007).
We estimate the mass-loss rate from the planet's atmosphere by assuming that the
outflow forms a spherically symmetric envelope around the planet which is optically thick
in the C II resonance line covering about 8% of the stellar surface. The spherical symmetry
assumption is a good approximation because the shape of the tidally induced Roche lobe
as seen during transit is close to a sphere (Lecavelier des Etangs et al. 2004). The envelope
becomes optically thin at its "limb" because gas expansion and mass flux conservation reduce
-- 11 --
the gas density with increasing radial distance (r) from the planet. We define the envelope
"limb" to be at the radial distance from the planet for which the absorption area is the same
as that of an opaque sphere of the same size.
We consider a line of sight from the star that passes through the envelope with p the
point of closest approach to the planet. The quantity x measures the distance along this
line of sight from point p back to the star. The optical depth to the observer in the C II
resonance line along this line of sight at the apparent limb of the envelope is,
τ (p, ∆ν) = 2kZ ∞
0
nC II(x)dx = 0.69,
(1)
√πe2
mc
f
∆νD
lines of sight passing outside of p, R ∞
where nC II is the number density of C+ ions, k =
is the line opacity, f = 0.128
is the oscillator strength, m is the electron mass, and ∆νD is the Doppler width. The value
of p depends somewhat on the displacement from line center, ∆ν, because for expanding
gas, most of the absorption occurs near the frequency corresponding to the line-of-sight
component of the expansion velocity. The limb occurs at τ = 0.69, where as much stellar
flux is transmitted through the envelope for lines of sight passing inside of p as for those
e−tdt, where t is the optical depth
through the envelope for lines of sight with different values of p. Since the gas density and
thus optical depth increase along lines of sight inside of p, the limb will occur at somewhat
smaller optical depths depending on details the density and outflow velocity. For example,
Lecavelier des Etangs et al. (2008a) calculated τ = 0.56 for an atmosphere in hydrostatic
equilibrium, which has a density that decreases exponentially outward but does not have an
outflow or heating in the outer layers. If the density decreases with radial position according
to a power law, then τ could be smaller than 0.69 by a factor of 2 or 3. The resulting mass
loss rates would also be smaller by this factor. We call attention to this uncertainty that
should be with resolved by realistic models based on physical principles that predict transit
line profiles consistent with the COS data.
0.69 e−tdt = R 0.69
0
Since the gas density decreases with increasing radial distance from the planet r (thus
increasing x along the line of sight), p should decrease slightly with increasing ∆ν. We
compute ∆νD = 2.7 × 1010 s−1, assuming a gas temperature of 10,000 K (Garcia Munoz
2007; Ballester et al. 2007). If the outflow is turbulent or some absorbers have superthermal
velocities (see below), then ∆νD will be larger. For a spherical outflow with constant mass
flux, the mass flux of C+ ions from the planet is
MC II = 4πr2mCvnC II(r),
(2)
-- 12 --
where v is the outflow speed at r, mC is the mass of a carbon atom, and r2 = x2 + p2.
We estimate MC II by requiring that the gas in the exosphere obscure 8% of the stellar
disk within ∆νD of line center, corresponding to ±3.7 km s−1. This condition will occur
when the optical depth is about 0.69 at p2 = 0.08R2
⋆, i.e.,
τ (p = 0.284R⋆, ∆ν) =
2√πe2f MC II
mc∆νD4πmCv Z ∞
0
dx
r2 = 0.69.
The integral has an analytical solution by setting u = x/p of
Z ∞
0
dx
r2 =
1
p Z ∞
0
du
u2 + 1
=
π
2p
.
(3)
(4)
We adopt v = 10 km s−1 at p = 0.284R⋆ = 2.36Rplanet from the theoretical models of
Tian et al. (2005), Garcia Munoz (2007), and Murray-Clay et al. (2009). We then find that
MC II = 2.1 × 107g s−1 = 6.8 × 1014g yr−1.
(5)
In his model for the exosphere of HD 209458b, Garcia Munoz (2007) found that essentially all
of the carbon is singly ionized. If we assume that the carbon abundance is solar, 2.7 × 10−4
that of hydrogen (Asplund et al. 2009), the total mass-loss rate from the planet for only
thermal broadening of the C II lines is
Mtotal = 8.0 × 1010g s−1 = 2.4 × 1018g yr−1.
(6)
Since the planet's mass is 1.2 × 1030 g (Knutson et al. 2007), the fractional mass-loss
rate per year is very small, as shown by
Mtotal
Mplanet
= 2.0 × 10−12yr−1.
(7)
This calculation assumes that the mass-loss rate and planetary orbit do not change with
time. Since additional line broadening is required to explain the broad absorption in the
MCII and Mtotal are at the lower end of their realistic ranges (see
C II lines, these values of
Section 5).
-- 13 --
5. DISCUSSION
We now compare our empirical mass-loss rate with those of various theoretical models.
All theoretical models of which we are aware assume hydrodynamical outflow driven by high
temperatures and pressures near the base of the outflow. The important questions are the
amount and location of energy absorbed from the star, the important cooling mechanisms,
and the location of the sonic point in these transonic (Parker-like) winds relative to the
location of the exoplanet's Roche lobe. Vidal-Madjar et al. (2003) demonstrated that neu-
tral hydrogen must be escaping from a volume larger than the planet's Roche lobe by an
unspecified mass-loss mechanism and that stellar radiation pressure further accelerates the
M > 1010
outflow and shapes a comet-like tail. They proposed that the mass-loss rate is
g s−1, but they added that "owing to saturation effects in the [Lyman-α] absorption line, a
flux larger by several orders of magnitude would produce a similar absorption signature."
Murray-Clay et al. (2009) computed models with and without radiation pressure and con-
M, but it can change the location of the
cluded that radiation pressure does not increase
mass loss from the day side to the night side of the exoplanet. Tidal forces also do not
change M appreciably, but they move the sonic point closer to the exoplanet and increase
the flow velocity (Garcia Munoz 2007; Murray-Clay et al. 2009).
All of the theoretical models assume that HD 209458 is a main sequence star like the Sun
with ultraviolet and extreme ultraviolet radiative output similar to the Sun at low or high
activity. Since mass-loss rates are proportional to the energy input (the so-called energy-
In all of the models,
limited escape rate), the critical question is the energy input rate.
expansion (PdV work) is the dominant cooling mechanism, although radiation by H+
3 at the
base of the flow must be included (Yelle 2004), and radiation in the Lyman-α line can be
important for T Tauri star radiative input rates. Many models assume that only the energy
remaining after stellar Lyman continuum (E> 13.6 eV) photons ionize neutral hydrogen is
available to energize the outflow, so that the assumed fraction of this radiation (83% or 100%
M is a few times 1010 g s−1. For
in different models) controls the outflow. For such models,
M = 4.7 × 1010, Tian et al. (2005)
example, Yelle (2004), as corrected in Yelle (2006), finds
M = 6.1 × 1010 for EUV flux similar to the
find M < 6 × 1010, Garcia Munoz (2007) finds
M = 1.5 × 1011 for EUV flux like the active Sun. Murray-Clay et al.
low activity Sun but
(2009) find that
M ∼ 2 × 1010 g s−1.
The inclusion of atoms and molecules other than H, H+, and H2 can significantly change
the outflow, in particular by increasing the mass-loss rate, because more of the stellar UV
radiation is available to heat the outflow and drive the mass loss. For example, the Lyman-
α line contains more flux than the entire Lyman continuum (Ribas et al. 2005), and other
UV emission lines and continuum radiation in the wavelength range 912 -- 2000 A can be
-- 14 --
absorbed in the bound-free continua of atomic species (e.g., CNO) and molecules including
M increased
H2. When Garcia Munoz (2007) added solar abundances of CNO to his models,
by a factor of 4.5 (see his Table 6). Including partial absorption of the 1000 -- 2000 A stellar
M to 6.3 × 1011 g s−1. Metals
flux by atoms and molecules with solar abundances increases
must exist in the distant outflow since we see absorption during transit in the C II and Si III
lines. By equating the energy input of both Lyman continuum and Lyman-α with stellar
fluxes estimated from the rotation rate of HD 209458, Lecavelier des Etangs (2007) found
M ∼ 3×1011 g s−1. Using a similar energy balance argument, Lammer et al. (2003) obtained
M ∼ 3 × 1011 g s−1, but Yelle (2004) suggested that this result could be an overestimate as
a result of not including H+
3 cooling and ionization of H.
A critical question is whether line absorption during transit is produced entirely by
the outflowing thermal plasma or whether high-velocity atoms and ions (either at higher
temperatures or nonthermal) are required to explain the broadest part of the absorption
line profiles. Ben-Jaffel & Sona Hosseini (2010) show that neutral hydrogen absorption by
scaled hydrodynamic outflow models may explain the flux decrease during transit even at
±100 km s−1 from the center of the Lyman-α line, but such models predict much less
absorption during transit than is detected in the unresolved O I and C II line flux measure-
ments (Vidal-Madjar et al. 2004). Additional high-velocity absorbers are needed to explain
the O I, C II, and probably also the Lyman-α data, but there is no agreement concern-
ing the source of the high-velocity absorbers. One possibility is radiation pressure, but
Murray-Clay et al. (2009) argue that the observed absorption near +100 km s−1 from the
center of the Lyman-α line is difficult to explain by stellar radiation pressure as originally pro-
posed by Vidal-Madjar et al. (2003). However, radiation pressure on neutral hydrogen atoms
must be present, and Lecavelier des Etangs et al. (2008b) argue that radiation pressure by it-
self can explain the Lyman-α observations. Holmstrom et al. (2008) and Murray-Clay et al.
(2009) argue that the observed decrease in Lyman-α flux near and beyond ±100 km s−1
during transit requires high-velocity H atoms that could be explained by charge exchange
between stellar wind protons and neutral H atoms in the exoplanet's outflow. These high-
velocity H atoms are called ENAs (energetic neutral atoms).
Ben-Jaffel & Sona Hosseini (2010) have tested whether the presence of high-velocity
atoms and ions could explain the flux reduction observed in the unresolved O I and C II
line transit data. They assume that a portion of the gas has an effective temperature
(either thermal or nonthermal) many times larger than the thermal temperature in outflow
models. For the unresolved C II 1334, 1335 A multiplet, they can explain the observed flux
reduction by C+ ions with effective temperatures 5.5 -- 107 times larger than the background
gas temperature (TB) in typical hydrodynamic models depending on the properties of the
outflow models and the assumed location of the C+ ions. In Figure 9 of their paper, they
-- 15 --
show examples of predicted C II 1334 A line profiles during transit with TCII/TB ∼ 16 and
∼ 28 that can explain the total decrease in unresolved line flux during transit, but these
particular examples are either too narrow or triangular in shape, unlike the observed C II
line profiles which show significant flux reduction between -- 50 and +50 km s−1 with probable
enhanced absorption near -- 10 and +15 km s−1. Ben-Jaffel (private communication) informs
us that other models with somewhat larger TCII/TB ratios can explain the flat reduction
in C II line flux between -- 50 and +50 km s−1. However these models have ad hoc velocity
distributions and they do not explain the observed velocity structure. In their paper, they
conclude that COS spectra are needed to "...reveal the true balance between thermal and
non-thermal populations...". We encourage the development of new models, preferably based
on plausible physical mechanisms, to explain the observed line profiles.
Our analysis of the C II line absorption during transit assuming only thermal absorption
Mtotal = 8.0× 1010 g s−1.
by 10,000 K gas and solar carbon abundance led to our estimate of
This mass-loss rate is somewhat larger than the theoretical estimates mentioned above that
assume an outflow with no metals. The assumption of no metals in the outflow is clearly
unrealistic as we observe absorption by C II and Si III at 2.36Rplanet. At the same time, our
analysis of the C II lines assuming only thermal line broadening (at 10,000 K) is unrealistic
as the thermal velocity is only 3.7 km s−1, and we observe absorption out to ±50 km s−1.
Since we measure the mass flux at the limb of the outflow where the C II optical depth
is small (τ = 0.69), we cannot invoke optical thickness to explain the broad profile. Note
that the same is true for the Lyman-α line, but (as the referee has called to our attention)
Ben-Jaffel & Sona Hosseini (2010) and other authors include opacity and Lorentzian line
shapes at the limb as a possible mechanism for absorption line broadening.
Equation 3 shows that MCII is proportional to the Doppler broadening parameter ∆νD.
To obtain 8% absorption at ±50 km s−1 from line center requires either turbulent broadening
with velocity of about 50 km s−1 or superthermal C+ ions with similar velocities. If turbulent
broadening is responsible, then the mass-loss rate increases by a factor of 13 to Mtotal =
1 × 1012 g s−1. However, 50 km s−1 turbulent velocities are highly supersonic and unlikely.
We estimate that a sensible upper limit to the turbulent velocity would be twice the flow
speed or 20 km s−1, for which Mtotal ≈ 4 × 1011 g s−1. This would increase Mtotal/Mplanet to
≈ 4 × 10−11 yr−1.
If high-velocity C+ ions are responsible for the broad absorption, then the mass-loss
rate could be closer to the thermal value because only a small fraction of the C+ ions are
Mtotal = (8 -- 40)×1010 g s−1 is consistent
likely to have high velocities. We conclude that
with the C II line profiles. We find it interesting that when Garcia Munoz (2007) includes
metals with solar abundances in his hydrodynamic mass outflow models, the mass-loss rate
-- 16 --
increases from M = 1.4 × 1011 to M = 4.95 × 1011 g s−1, a value near the top end of our
empirical mass-loss rate estimate. Thus high-velocity C+ ions are required to explain the C II
absorption during transit. Similar arguments would likely explain the broad Si III transit
absorption.
6. CONCLUSIONS
The high sensitivity and moderate spectral resolution of COS allowed us to obtain the
first measurements of gas absorption in the exosphere of a transiting exoplanet in lines of
C II and Si III with velocity resolution. During transits of HD 209458, we found that the
mean flux in these lines is reduced by 7.8 ± 1.3% for the co-added C II lines and 8.2 ± 1.4%
for the Si III line in the velocity range -- 50 to +50 km s−1. The 8% absorption constraint
sets the location of the outflow limb (2.36Rplanet) where the C II lines become optically thin.
Theoretical hydrodynamic models show that the outflow velocity is about 10 km s−1 and
gas temperature about 10,000 K at this location. If we assume that the outflow is purely
thermal, then Mtotal = 8 × 1010 g s−1, somewhat larger than theoretical estimates based on
outflows with no metals. Since absorption in the C II lines is much broader than indicated
by Doppler broadening with a thermal velocity of 3.7 km s−1, the lines must be broadened by
large turbulence or high-velocity C+ ions. If turbulent broadening dominates, then the 8%
absorption seen out to ±50 km s−1 would require mass-loss rates as large as 1 × 1012 g s−1,
but such highly supersonic velocities are not likely. Instead, absorption by high-velocity C+
and Si++ ions formed by the interaction of the stellar wind and planetary mass loss or some
other mechanism broadens the C II and Si III lines. We conclude that the mass-loss rate
in the exoplanet's outflow is in the range (8 -- 40)×1010 g s−1. These rates assume that the
carbon abundance is solar, which is not the case for the giant planets in the solar system.
We note that mass-loss rates predicted by hydrodynamic models that include solar abundant
metals in the outflow lie at the upper limit of this range.
Comparison of the C II and Si III line depths and transit/non-transit line ratios shows
deep absorption features near -- 10 and +15 km s−1 in both lines. Since the C II and Si III
lines are likely formed in the same region of the exosphere, we include in Figure 4 the co-
addition of the two profiles with typical error bars for comparison with theoretical models.
The increase in S/N with the co-addition of the C II and Si III spectra allows us to make
more definitive statements about the velocity structure. The dips near -- 10 and +15 km s−1
are more than twice as large as the errors per spectral resolution element (about 17 km s−1)
and are most likely real. In addition, the ratios near -- 40 and +30 -- 70 km s−1 are also low, but
the line fluxes decrease rapidly away from line center and these low ratios are less certain.
-- 17 --
Future observations with higher S/N are needed to confirm this first detection of velocity
structure in the expanding atmosphere of an exoplanet.
The COS transit observations of the absorption velocity structure in these lines demon-
strate the need for including high-velocity C+ and Si++ ions in models of the exosphere
of HD 209458b and presumably other close-in exoplanets. It remains to be seen whether
the inclusion of a simple distribution of ions with high effective temperature as proposed
by Ben-Jaffel & Sona Hosseini (2010) is adequate to explain the data. The new COS data
provide the opportunity to test a range of absorption models based on such physical mecha-
nisms as radiation pressure, charge exchange, and interactions between the stellar wind and
the planet's outflow to understand mass-loss processes in the exospheres of close-in planets.
This work is supported by NASA through grants NNX08AC146 and NAS5-98043 to
the University of Colorado at Boulder. We thank the many people at the Space Telescope
Science Institute, Goddard Space Flight Center, Ball Aerospace and Technologies Corp. and
CASA at the University of Colorado for the excellent hardware and software that has made
COS a great success. We particularly wish to thank the referee for his many thoughtful and
thorough comments and his bringing to our attention important publications that have led
to significant modifications in the paper. We thank Dr. J. Schneider for encouraging us to
look for an assymetry between ingress and egress. We also thank Dr. Lofti Ben-Jaffel for
clarifying our understanding of the Ben-Jaffel & Hosseini (2010) paper. We thank Dr. Alain
Lecavelier for calling our attention to the different values of τ at the limb that depend on
the density structure in the envelope.
REFERENCES
Asplund, M., Grevesse, N., Sauval, J., & Scott, P. 2009, ARA&A, 47, 481
Baliunas, S. L. & Butler, S. E. 1980, ApJ, 235, L45
Ballester, G. E., Sing, D. K., & Herbert, F. 2007, Nature, 445, 511
Baraffe, I., Selsis, F., Chabrier, G., Barman, T. S., Allard, F., Hauschildt, P. H., & Lammer,
H. 2004, A&A, 419, L13
Ben-Jaffel, L. 2007, ApJ, 671, L61
Ben-Jaffel, L. 2008, ApJ, 688, 1352
Ben-Jaffel, L. & Sona Hosseini, S. 2010, ApJ, 709, 1284
-- 18 --
Charbonneau, D., Brown, T. M., Latham, D., & Mayor, M. 2000, ApJ, 529, L45
Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L. 2002, ApJ, 568, 377
Ehrenreich, D. et al. 2008, A&A, 483, 933
Garcia Munoz, A. 2007, Planet. Space Sci., 55, 1426
France, K., Stocke, J. T., Yang, H., Linsky, J. L., Froning, C. S., Green, J. C., & Osterman,
S. N. 2010, ApJ, to appear
Ghavamian, P. et al. 2010 BAAS, 42, 499
Henry, G. W., Marcy, G. W., Butler, R. P., & Vogt, S. S. 2000, ApJ, 529, L41
Holmstrom, M., Ekenback, A., Selsis, F., Penz, T., Lammer, H., & Wurz, P. 2008, Nature,
451, 970
Jaritz, G. F., Endler, S., Langmayr, D., Lammer, H., Griessmeier, J.-M., Erkaev, N. V., &
Biernat, H. K. 2005, A&A, 439, 771
Koskinen, T. T., Aylward, A. D., & Miller, S. 2007, Nature, 450, 845
Koskinen, T. T., Yelle, R. V., Lavvas, P., & Lewis, N. K. 2010, arXiv:1004.1396
Knutson, H. A., Charbonneau, D., Noyes, R. W., Brown, T.M., & Gilliland, R. L. 2007,
ApJ, 655, 564
Lammer, H., Selsis, F., Ribas, I., Guinan, E. F., Bauer, S. J., & Weiss, W. W. 2003, ApJ,
598, L121
Lecavelier des Etangs, A. 2007, A&A, 461, 1185
Lecavelier des Etangs, A., Vidal-Madjar, A., McConnell, J. C., & H´ebrard, G. 2004, A&A,
418, L1
Lecavelier des Etangs, A., Pont, F., Vidal-Madjar, A., & Sing, D. 2008a, A&A, 481, L83
Lecavelier des Etangs, A., Vidal-Madjar, A., & Desert, J.-M. 2008b, Nature, 458, E1
Mazeh, T. et al. 2000, ApJ, 532, L55
Murray-Clay, R. A., Chiang, E. I., & Murray, N. 2009, ApJ, 693, 23
Redfield, S. & Linsky, J. L. 2008, ApJ, 673, 283
-- 19 --
Ribas, I., Guinan, E. F., Gudel, M., & Audard, M. 2005, ApJ, 622, 680
Rottman, G. 2006, Space Sci. Rev., 125, 39
Schneider, J., Rauer, H., Lasota, J. P., Bonazzola, S., & Chassefi`ere, E. 1998, in ASP Conf.
Series 134, Brown Dwarfs and Extrasolar Planets, ed. R. Rebolo, E. L. Mart´ın, & M.
R. Zapatero Osorio (San Francisco, ASP), 241
Schneiter, E. M., Vel´azquez, P. F., Esquivel, A., Raga, A. C., & Blanco-Cano, X. 2007 ApJ,
671, 57
Tian, F., Toon, O. B., Pavlov, A. A., & De Sterck, H. 2005, ApJ, 621, 1049
Wittenmyer, R. A. et al. 2005, ApJ, 632, 1157
Vidal-Madjar, A. et al. 2003, Nature, 422, 143
Vidal-Madjar, A. et al. 2004, ApJ, 604, L69
Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., Ballester, G. E., Ferlet, R.,
H´ebrard, G., & Mayor, M. 2008, ApJ, 676, L57
Yelle, R. V. 2004, Icarus, 170, 167
Yelle, R. V. 2006, Icarus, 183, 508
This preprint was prepared with the AAS LATEX macros v5.2.
-- 20 --
Transit
Non−Transit
Si IV
Cl I O I
O I
1360
1380
Wavelength (Å)
Si IV
O IV]
O IV]
1400
C II
C II
)
1
−
Å
2
−
m
c
1
−
s
g
r
e
5
1
−
0
1
(
x
u
F
l
5
4
3
2
1
0
1340
Fig. 1. -- A portion of the COS G130M spectrum of HD 209458. The lower (black) plot is
the sum of four spectra obtained during transit. The upper (red) plot (displaced upward) is
the sum of the ten quadrature and four secondary eclipse spectra.
-- 21 --
Si III 1206.50 Å
Si IV 1393.76 Å
8
6
4
2
0
−100
0.4
0.0
)
1
−
Å
2
−
m
c
1
−
s
g
r
e
5
1
−
0
1
(
x
u
F
l
−0.4
−0.8
−1.2
−100
−50
1.2
1.0
0.8
0.6
o
i
t
a
R
Transit
coaddition of 4 exposures
Non−transit
coaddition of 14 exposures
−50
0
Velocity (km s−1)
Si III 1206.50 Å. Difference: black−red
50
100
Transit
coaddition of 4 exposures
Non−transit
coaddition of 14 exposures
−50
0
Velocity (km s−1)
Si IV 1393.76 Å. Difference: black−red
50
100
4
3
2
1
0
−100
0.4
0.0
)
1
−
Å
2
−
m
c
1
−
s
g
r
e
5
1
−
0
1
(
x
u
F
l
−0.4
0
Velocity (km s−1)
Si III 1206.50 Å. Ratio: black/red
50
100
−100
−50
0
Velocity (km s−1)
Si IV 1393.76 Å. Ratio: black/red
50
100
1.2
1.0
0.8
0.6
o
i
t
a
R
−100
−50
0
Velocity (km s−1)
50
100
0.4
−100
−50
0
Velocity (km s−1)
50
100
Fig. 2. -- Upper panels: Comparison of the four co-added spectra obtained during transit
(black) with the co-added non-transit spectra (red) for the Si III and Si IV lines. Repre-
sentative error bars per pixel are included. Middle panels: difference spectra (transit minus
non-transit) with representative error bars. Lower panels: flux ratio (transit/non-transit)
spectra. The ratios are not plotted beyond ±50 km s−1 as the errors are large at low flux
levels. The velocity scale is relative to the -- 14.8 km s−1 radial velocity of the star.
-- 22 --
C II 1334.53 Å
C II 1335.69 Å
Transit
coaddition of 4 exposures
Non−transit
coaddition of 14 exposures
ISM
−50
0
Velocity (km s−1)
50
100
C II 1334.53 Å. Difference: black−red
3
2
1
0
−100
0.4
0.0
)
1
−
Å
2
−
m
c
1
−
s
g
r
e
5
1
−
0
1
(
x
u
F
l
Transit
coaddition of 4 exposures
Non−transit
coaddition of 14 exposures
−50
0
Velocity (km s−1)
50
100
C II 1335.69 Å. Difference: black−red
6
5
4
3
2
1
0
−100
0.4
0.0
−0.4
−0.8
−1.2
Coaddition of both C II lines
Transit
Non−transit
−50
0
Velocity (km s−1)
Difference: black−red
50
100
8
7
6
5
4
3
2
1
0
−100
0.4
0.0
−0.4
−0.8
−1.2
−0.4
−0.8
−100
1.2
1.0
0.8
0.6
o
i
t
a
R
−50
0
Velocity (km s−1)
50
100
−100
−50
0
Velocity (km s−1)
50
100
−100
−50
C II 1334.53 Å. Ratio: black/red
C II 1335.69 Å. Ratio: black/red
1.2
1.0
0.8
0.6
1.2
1.0
0.8
0.6
0
Velocity (km s−1)
Ratio: black/red
50
100
−100
−50
0
Velocity (km s−1)
50
100
−100
−50
0
Velocity (km s−1)
50
100
−100
−50
0
Velocity (km s−1)
50
100
Fig. 3. -- Upper panels: Comparison of the four co-added spectra obtained during transit
(black) with the co-added non-transit spectra (red) for the C II 1334.53 A and 1335.69 A
lines. The upper right panel show a co-addition of both C II lines on the same velocity
scale. Representative error bars per pixel are included. Middle panels: difference spectra
(transit minus non-transit) with representative error bars for each C II line and the combined
C II line. Lower panels: flux ratio (transit/non-transit) spectra. The ratios are not plotted
beyond ±50 km s−1 as the errors are large at low flux levels. The velocity scale is relative
to the -- 14.8 km s−1 radial velocity of the star.
0.4
0.0
−0.4
−0.8
−1.2
−100
−50
C II lines coaddition
Si III 1206 line
−50
1.2
1.0
0.8
0.6
0.4
−100
1.2
1.0
0.8
0.6
0.4
−100
-- 23 --
Difference
0
Ratio
0
Ratio
C II lines coaddition
Si III 1206 line
50
100
50
100
Coaddition of C II and Si III lines
−50
0
Velocity (km s−1)
50
100
Fig. 4. -- Upper and middle panels: Comparison of the difference spectra (transit minus non-
transit) and ratio spectra (transit/non-transit) for the co-addition of the C II Lines (black)
and the Si III line (red). Lower panel: Co-addition of the C II and Si III ratio spectra with
typical errors per spectral resolution element (about 17 km s−1). The horizontal line at 0.92
is the mean ratio for the velocity interval -- 50 to +50 km s−1. The dips near -- 10 and +15
km s−1 are each more than 2σ and likely real. The low ratio features near -- 40 and +30 -- 70
km s−1 occur where the line fluxes are low and are thus less certain.
-- 24 --
Table 1. Propertiesa of HD 209458 and HD 209458b
Spectral type
Distance (pc)
M⋆/M⊙
R⋆/R⊙
Porbit (days)
Rplanet/RJup
Mplanet/MJup
ρplanet/ρJup
Semimajor axis (AU)
Stellar flux at planet/Jupiter
Escape speed (km/s)
Orbital Velocity ampl. (km/s)
Transit duration (hr)
Transit depth for Rplanet
Transit depth for RRoche
G0 V
47
1.101+0.066
−0.062
1.125+0.020
−0.023
3.52474859(38)
1.320+0.024
−0.025
0.64 ± 0.06
0.26 ± 0.04
0.045
13,400
42 ± 4
146
3.22
1.5%
10%
aData from Knutson et al. (2007).
-- 25 --
Table 2. COS G130M Observations of HD 209458 (Program 11534)
Exp. Time (s)
Phase
Day
Start
Stop
2340.2
0955.2
1851.2
0560.0
2235.1
2340.2
0945.2
1787.2
0925.2
1798.2
1045.2
1096.2
1400.2
1405.2
1057.2
1093.1
1401.2
1405.2
0.2516 -- 0.2593
0.2684 -- 0.2716
0.2722 -- 0.2782
0.2873 -- 0.2891
0.2898 -- 0.2971
0.7059 -- 0.7136
0.7229 -- 0.7260
0.7266 -- 0.7324
0.7417 -- 0.7448
0.7454 -- 0.7513
0.9913 -- 0.9947
0.9954 -- 0.9990
0.0077 -- 0.0123
0.0129 -- 0.0175
0.4880 -- 0.4915
0.4921 -- 0.4957
0.5043 -- 0.5089
0.5095 -- 0.5141
19 Sep
19 Sep
19 Sep
19 Sep
19 Sep
24 Sep
24 Sep
24 Sep
24 Sep
24 Sep
02 Oct
02 Oct
02 Oct
02 Oct
18 Oct
18 Oct
18 Oct
18 Oct
10:10:15
11:35:37
11:54:40
13:11:29
13:24:06
13:12:09
14:38:04
14:56:57
16:13:54
16:32:36
14:32:10
14:52:52
15:55:28
16:21:56
10:57:04
11:17:49
12:19:37
12:46:06
10:49:15
11:51:32
12:25:31
13:20:49
14:01:21
13:51:09
14:53:49
15:26:44
16:29:19
17:02:34
14:49:35
15:11:08
16:18:48
16:45:21
11:14:41
11:36.02
12:42:58
13:09:31
Gapa (A)
1278 -- 1290
1288 -- 1299
1296 -- 1306
1288 -- 1299
1306 -- 1316
1278 -- 1290
1288 -- 1299
1296 -- 1306
1288 -- 1299
1306 -- 1316
1278 -- 1290
1288 -- 1299
1296 -- 1306
1306 -- 1316
1278 -- 1290
1288 -- 1299
1296 -- 1306
1306 -- 1316
aThe gap is the wavelength interval between the two detector faceplate el-
ements. No data are available for the gap region, and there are instrumental
artifacts at wavelengths close to the gap.
|
1508.04795 | 1 | 1508 | 2015-08-19T21:00:28 | Jupiter's Deep Cloud Structure Revealed Using Keck Observations of Spectrally Resolved Line Shapes | [
"astro-ph.EP"
] | Technique: We present a method to determine the pressure at which significant cloud opacity is present between 2 and 6 bars on Jupiter. We use: a) the strength of a Fraunhofer absorption line in a zone to determine the ratio of reflected sunlight to thermal emission, and b) pressure-broadened line profiles of deuterated methane (CH3D) at 4.66 microns to determine the location of clouds. We use radiative transfer models to constrain the altitude region of both the solar and thermal components of Jupiter's 5-micron spectrum. Results: For nearly all latitudes on Jupiter the thermal component is large enough to constrain the deep cloud structure even when upper clouds are present. We find that Hot Spots, belts, and high latitudes have broader line profiles than do zones. Radiative transfer models show that Hot Spots in the North and South Equatorial Belts (NEB, SEB) typically do not have opaque clouds at pressures greater than 2 bars. The South Tropical Zone (STZ) at 32 degrees S has an opaque cloud top between 4 and 5 bars. From thermochemical models this must be a water cloud. We measured the variation of the equivalent width of CH3D with latitude for comparison with Jupiter's belt-zone structure. We also constrained the vertical profile of water in an SEB Hot Spot and in the STZ. The Hot Spot is very dry for P<4.5 bars and then follows the water profile observed by the Galileo Probe. The STZ has a saturated water profile above its cloud top between 4 and 5 bars. | astro-ph.EP | astro-ph |
Jupiter's Deep Cloud Structure Revealed Using Keck
Observations of Spectrally Resolved Line Shapes
G. L. Bjoraker
NASA/GSFC Code 693, Greenbelt, MD 20771, USA
[email protected]
Department of Astronomy, University of California, Berkeley, CA 94720-3411, USA
M. H. Wong
Department of Astronomy, University of California, Berkeley, CA 94720-3411, USA
I. de Pater
and
M. ´Ad´amkovics
Department of Astronomy, University of California, Berkeley, CA 94720-3411, USA
ABSTRACT
Technique: We present a method to determine the pressure at which significant cloud opacity
is present between 2 and 6 bars on Jupiter. We use: a) the strength of a Fraunhofer absorption
line in a zone to determine the ratio of reflected sunlight to thermal emission, and b) pressure-
broadened line profiles of deuterated methane (CH3D) at 4.66 µm to determine the location of
clouds. We use radiative transfer models to constrain the altitude region of both the solar and
thermal components of Jupiter's 5-µm spectrum. Results: For nearly all latitudes on Jupiter the
thermal component is large enough to constrain the deep cloud structure even when upper clouds
are present. We find that Hot Spots, belts, and high latitudes have broader line profiles than do
zones. Radiative transfer models show that Hot Spots in the North and South Equatorial Belts
(NEB, SEB) typically do not have opaque clouds at pressures greater than 2 bars. The South
Tropical Zone (STZ) at 32◦S has an opaque cloud top between 4 and 5 bars. From thermochemical
models this must be a water cloud. We measured the variation of the equivalent width of
CH3D with latitude for comparison with Jupiter's belt-zone structure. We also constrained
the vertical profile of H2O in an SEB Hot Spot and in the STZ. The Hot Spot is very dry for
P<4.5 bars and then follows the H2O profile observed by the Galileo Probe. The STZ has a
saturated H2O profile above its cloud top between 4 and 5 bars.
Subject headings: planets and satellites: individual (Jupiter) -- planets and satellites: atmospheres
1.
INTRODUCTION
The abundance of H2O in Jupiter's atmosphere
is of fundamental importance in understanding the
origin of Jupiter, the composition of its clouds,
and Jovian dynamics beneath the upper cloud
layers. Water was first detected on Jupiter by
Larson et al. (1975) using the Kuiper Airborne
Observatory (KAO). Bjoraker et al. (1986a,b) an-
alyzed both KAO and Voyager IRIS spectra in
Jupiter's 5-µm spectral window to the deep at-
1
mosphere, finding a deep water abundance highly
depleted with respect to the solar abundance.
The spectrum between 4.5 and 5.4 µm provides
a wealth of information about the gas composi-
tion and cloud structure of the troposphere of this
giant planet. Jupiter's 5-µm spectrum is a mix-
ture of scattered sunlight and thermal emission
that changes significantly between belts and zones.
Jupiter exhibits remarkable spatial structure at
5 µm. Chemical models of Jupiter's cloud struc-
ture predict three distinct layers: an NH3 ice cloud
near 0.5 bars, an NH4SH cloud formed from a re-
action of NH3 and H2S at 2 bars, and a massive
water ice/liquid solution cloud near 5 or 6 bars, de-
pending on assumptions of composition and ther-
mal structure (see Weidenschilling & Lewis (1973)
and Wong et al. (2015)). Thermal emission from
the deep atmosphere is attenuated by the variable
opacity of one or more of these three cloud lay-
ers. Hot Spots, located primarily in the North and
South Equatorial Belts (NEB, SEB), exhibit 5-
µm radiances up to 70 times larger than surround-
ing regions due to a minimum of cloud opacity.
They also appear brighter than their surroundings
at microwave wavelengths due to a low ammonia
abundance (Sault et al. 2004).
The interpretation of 5-µm spectra of Jupiter,
and the H2O abundance in particular, has been
hampered by uncertainties in the pressure level of
the lower boundary of the emitting region. Two
different models have been proposed. Bjoraker et al.
(1986a,b) suggested that thermal radiation at
5 µm originates from levels as deep as 8 bars,
310 K where unit optical depth in H2 occurs. In
contrast, Carlson et al. (1992) proposed a model
in which a massive water-ice cloud establishes
the lower boundary near 5 bars, 273 K. Bjoraker
et al. fitted KAO and Voyager IRIS spectra of
H2O in Jupiter's Hot Spots with a small abun-
dance (4-30 ppm) distributed along a long path
(60 km) between 2 and 8 bars. They also mea-
sured H2O abundances in the low-flux zone regions
using Voyager spectra. The NEB hot spots were
found to be depleted in H2O between 2 and 4 bars,
but belts, zones, and hot spots could all be fitted
by the same H2O profile (4-30 ppm) between 4
and 8 bars. In marked contrast, Carlson et al. fit-
ted Voyager Hot-Spot spectra using a much larger
mixing ratio (up to 3000 ppm, equivalent to 3×
the solar O/H measured by Asplund et al. (2009))
2
confined to a narrow layer (10 km) between 4 and
5 bars immediately above an opaque water cloud.
Carlson et al. also examined Voyager spectra of
the Equatorial Zone (EQZ) and other regions away
from hot spots. All regions required a saturated
H2O profile that increased from 300 ppm at 4
bars to 3000 ppm at 5 bars, although NEB hot
spots were sub-saturated in the 2 to 4 bar region.
The Galileo Probe measured water vapor in
situ at a single location on Jupiter, but unan-
swered questions remain about its global abun-
dance. The probe entered Jupiter's atmosphere
at 6.5◦N planetocentric latitude near the south-
ern portion of an NEB hot spot. The Galileo
Probe Mass Spectrometer (Niemann et al. 1992)
found water increasing with depth, from an up-
per limit of 0.8 ppm at 2.7 bar, to measured
mole fractions of 40±13 ppm at 11.0-11.7 bar and
420±140 ppm at 18-21 bar (Niemann et al. (1998)
and Wong et al. (2004b)). The deepest value cor-
responds to 0.45× solar O/H. Ground-based imag-
ing has shown that 5-µm hot spots, such as the
one entered by the Galileo probe, cover less than
1% of the surface area of Jupiter (Orton et al.
1996). Thus, the H2O abundances observed by
the probe may be characteristic of all or most hot
spots, but they are probably not representative
of Jupiter as a whole, especially since a range of
other indirect studies (lightning flash depths, the
tropospheric CO abundance, and discrete clouds
at P ≥ 4 bar) point to solar or supersolar water
abundances (Wong et al. 2008). Since the Galileo
probe found that carbon, sulfur, and nitrogen were
enriched by ∼4× solar (Wong et al. 2004b), oxy-
gen may also be enhanced by the same amount.
Some formation models require ∼10× solar O/H
in order to trap Jupiter's volatiles inside cages of
water ice clathrates (Hersant et al. 2004). The
Juno mission, scheduled to begin orbiting Jupiter
in July 2016, should answer many of these ques-
tions. The Microwave Radiometer will measure
water vapor below Jupiter's clouds to determine
the O/H ratio (Janssen et al. 2005).
Interpreta-
tion of these data may not be straightforward,
however, due to e.g., the small microwave ab-
sorptivity of H2O gas compared with NH3
(see
de Pater et al. (2005) for details). A complemen-
tary way to determine the deep H2O abundance,
therefore,
is highly desirable. The Jovian In-
frared Auroral Mapper (JIRAM) will acquire near-
infrared spectra of Jupiter, including the 5-micron
window (Adriani et al. 2014).
(2001)
Water ice has been detected in isolated regions
where active convection lofted the ice well above
its condensation level (Simon-Miller et al. 2000),
but water clouds are generally hidden by overlying
NH3 and NH4SH clouds (e.g. Sromovsky & Fry
(2010)). Only very thin clouds were found in
the Galileo Probe Hot Spot (Ragent et al. 1998),
consistent with condensation via weak turbu-
lent updrafts within the descending branch of
an equatorially-trapped Rossby wave (Friedson
(2005), Wong et al. (2015)). Models of NIMS
spectra by Nixon et al.
include water
clouds in at least some Hot Spots, and Roos-Serote et al.
(2004) showed that NIMS spectra of Hot Spots
cannot rule out water clouds whose opacity is
entirely restricted to P > 5 bar. Evidence for
water clouds in Jupiter's zones is also ambigu-
ous. Drossart et al. (1998) compared dayside and
nightside NIMS spectra of low-flux regions in the
EQZ. A saturated H2O profile above an opaque
water cloud at 5 bars provides a satisfactory fit
to these data. At the spectral resolution of IRIS
(4.3 cm−1) and NIMS (10 cm−1) we can only
retrieve a column abundance of H2O above an
assumed lower boundary, which can be either a
water cloud or opacity due to H2. Thus, we sim-
ply cannot tell whether or not Hot Spots or zones
have water clouds. This problem will also affect
the interpretation of 5-µm spectra from JIRAM
on Juno, which has the same spectral resolution
as Voyager/IRIS (Adriani et al. 2014).
In Section 2 we present ground-based observa-
tions of Jupiter's belts and zones that have suffi-
cient spectral resolution to resolve line shapes. We
demonstrate that we can derive the cloud struc-
ture at P > 2 bars even when higher-altitude
clouds greatly attenuate the thermal flux from
the deep atmosphere. We used line shapes to
detect water clouds and to determine the pres-
sure at which these clouds become optically thick.
This resolves the ambiguity of whether the lower
boundary of the 5-µm line formation region is due
to H2 or due to opaque clouds. This, in turn, will
yield more accurate gas abundances for use in con-
straining models of Jupiter's origin and in under-
standing the dynamics of the atmosphere beneath
the upper clouds.
2. OBSERVATIONS
Five-micron spectra of Jupiter were acquired
using NIRSPEC on the Keck 2 telescope on March
11, 2014. NIRSPEC is an echelle spectrograph
with 3 orders dispersed onto a 1024x1024 InSb ar-
ray at 5 µm at our selected grating/cross-disperser
settings of 60.48 / 36.9 (McLean et al. 1998). A
0.4′′× 24′′slit was aligned north-south on the cen-
tral meridian of Jupiter, resulting in spectra with
a resolving power of 20,000. NIRSPEC has an ad-
vantage over instrumentation on other telescopes
(for example CSHELL on the IRTF) due to the
fact that 3 echelle orders at 5 µm are placed on
the detector array. Thus, each pixel along the slit
corresponding to different latitudes on Jupiter has
simultaneous spectra at 4.6, 5.0, and 5.3 µm. In
this paper we will focus on Order 16 covering 2131-
2165 cm−1 (4.62-4.69 µm) and Order 15 which
covers 1999-2031 cm−1 (4.92-5.00 µm). Jupiter
subtended 41′′ and the geocentric Doppler shift
was 26.3 km/sec. The water vapor column above
Mauna Kea was 2 precipitable mm derived from
fitting telluric lines in the Jupiter spectra. We did
not use stellar spectra for flux calibration or at-
mospheric transmission because the humidity dou-
bled between the time of the Jupiter and stellar
observations. The flux calibration is described in
Section 3. Fig. 1 shows an image of Jupiter us-
ing the SCAM guide camera on NIRSPEC. Al-
though the spectroscopy was performed at 5-µm,
the guide camera works at shorter wavelengths. A
K-prime filter centered at 2.12 µm was used to
obtain sufficient contrast to separate the bands of
variable haze reflectivity overlying Jupiter's belts
and zones. Two slit positions were required to ob-
tain pole-pole spectra of Jupiter. In Fig. 1 spatial
pixels that exhibit maxima in flux at 4.66 µm are
shown in red, locations that exhibit narrow line
profiles at 4.66 µm are shown in blue, and regions
that have broad pressure-broadened line profiles
are shown in green. We focus on two spatial loca-
tions with characteristically different spectra: Re-
gion A in the South Tropical Zone (STZ) at 32◦S,
and Region B, a Hot Spot in the SEB at 17◦S. We
developed radiative transfer models to calculate
synthetic spectra for regions A and B, as described
below.
The band pass at 4.66 µm contains absorption
lines of deuterated methane (CH3D), phosphine
3
Variation in CH3D Line Profiles
Equivalent Width Integration
1.0
0.8
Solar
0.6
0.4
0.2
0.0
PH3
EQZ
NTZ
56oN
52oN
NEB Hot Spot
CH3D
H2O
T
i
e
c
n
a
d
a
R
d
e
z
i
l
a
m
r
o
N
2142.0
2143.0
2144.0
2145.0
2146.0
Wavenumber
Fig. 2. -- Jupiter spectra at 4.66 µm (NIRSPEC order
16) show dramatic variation with latitude in the strength
and width of CH3D lines. Four absorption features are de-
noted by vertical lines. Narrow CH3D lines are observed in
the Equatorial and North Tropical Zones (EQZ, NTZ), but
they are much broader in Hot Spots and at high northern
latitudes. A phosphine (PH3) line and a Doppler-shifted
H2O line are shown. Fraunhofer lines (CO in the Sun) are
only observed in low-flux zones. T denotes telluric lines.
NIRSPEC order 16) of 5 regions in Jupiter's north-
ern hemisphere that exhibit either a maximum
in flux (an NEB Hot Spot at 8.5◦N), minima
in the equivalent width of CH3D (the EQZ at
0.5◦N and NTZ at 23◦N), or local maxima in
CH3D equivalent width. A horizontal bar indi-
cates the limits that we chose for numerical in-
tegration of the equivalent width that includes
six absorption lines of CH3D (blended into four
at this spectral resolution). All spectra are nor-
malized to 1.0 at 2141.6 cm−1 to facilitate com-
parison of line shapes, since the radiance of the
NEB Hot Spot is 70 times that of the NTZ. A tel-
luric H2O line, its Doppler-shifted counterpart,
a PH3 feature, and a Fraunhofer line due to CO
in the Sun are also shown. Note that the Fraun-
hofer line is observed only in low-flux zone regions
such as the EQZ and NTZ. In Sections 3 and 4
we describe how the strength of this feature can
be used to constrain cloud models. The equivalent
width of the set of CH3D lines at 2144 cm−1 varies
dramatically from belt to zone. Hot Spots and
belts exhibit broad pressure-broadened line pro-
files, while zones have much narrower features
that are resolved as 4 distinct CH3D absorption
lines. Note that an isolated CH3D line would be
spectrally resolved at a resolving power of 20,000
SEB Hot Spot
STZ
B
A
Fig. 1. -- Keck/SCAM image of Jupiter at 2.12 µm shows
2 NIRSPEC slit positions, one in the northern, and one in
the southern hemisphere. Curvature is due to navigating
positions onto an image taken a few minutes earlier. Red
pixels denote Hot Spots. Blue pixels denote positions of
minima in deuterated methane (CH3D) equivalent width
and thus candidates for opaque clouds. Green regions de-
note maxima in CH3D equivalent width and thus regions
without opaque clouds. Radiative transfer models were
used to calculate synthetic spectra for the zone marked A
and the SEB Hot Spot labeled B.
(PH3), and H2O. Methane and its isotopologues
do not condense, and they are not destroyed pho-
tochemically, in Jupiter's troposphere. We there-
fore assume that CH4 and CH3D have a con-
stant mixing ratio with respect to H2 in Jupiter's
troposphere, which means that variations in the
strength and shape of CH4 and CH3D lines be-
tween belts and zones on Jupiter should be due to
changes in cloud structure, not gas concentration.
Bjoraker et al. (2002) measured the spatial varia-
tion of the weak ν 3- ν 4 band of CH4 at 5.18 µm in
an attempt to derive cloud structure. Unfortu-
nately, they were unable to calculate synthetic
spectra that fit this feature. This is possibly due to
inaccurate or incomplete spectroscopic parameters
such as line strengths and broadening parameters.
Spectroscopic parameters for the much stronger
ν 2 fundamental band of CH3D at 2200 cm−1, or
4.5 µm, are well known (Nikitin et al. 1997) and
as shown below we now are able to derive cloud
structure and spatial variations therein.
Fig. 2 shows spectra at 4.66 µm (2144 cm−1,
4
(0.11 cm−1 resolution). This is because molecules
such as CH3D typically have broadening coeffi-
cients of 0.06 cm−1/atm and the line formation
region on Jupiter at 4.66 µm takes place at pres-
sures greater than 2 bars.
3. DATA ANALYSIS AND RESULTS
Absolute flux calibration was performed for
both NIRSPEC orders 15 and 16 by smoothing
the spectrum of the NEB Hot Spot at 8.5◦N to
4.3 cm−1 resolution, dividing by the transmittance
of the Earth's atmosphere above Mauna Kea, and
scaling the radiance of the resulting spectrum to
an average of the 4 hottest spectra of Jupiter's
NEB observed by Voyager IRIS in 1979.
In addition to pressure-broadened line pro-
files, the spectrum at 4.66 µm provides valu-
able information on the ratio of reflected sun-
light to thermal emission on Jupiter, which is
critical for understanding Jupiter's cloud struc-
ture. We compared the equivalent width of the
Fraunhofer line at 2141.8 cm−1 with its measured
value in the Sun using data from the Atmospheric
Trace Molecule Spectroscopy (ATMOS) experi-
ment which flew on the space shuttle Challenger
in 1985 (Farmer & Norton 1989; Farmer 1994).
Solar Fraunhofer lines in spectra of Jupiter arise
from regions with a significant fraction of reflected
sunlight, such as locations with thick high clouds.
Fraunhofer lines are not observed in regions that
are dominated by thermal emission, such as Hot
Spots. We selected a zone at 32◦S (marked"A"
on Figures 1, 4, and 5) for further study based
on the strength of this particular Fraunhofer line.
It is 43% as strong as in the Sun; thus, 57% of
the flux consists of thermal emission originating
in the deep atmosphere that has been attenu-
ated by one or more cloud layers before escaping
to space. This thermal flux preserves the broad
line profiles of CH3D caused by collisions with
H2
and He in its line formation region at the
deepest levels probed. We excluded the spectrum
of the NTZ shown in Fig. 2 for studies of the deep
atmosphere because the Fraunhofer line is 95%
as strong as in the Sun; thus, there is insufficient
thermal flux to constrain the deep cloud structure
at 23◦N. We also modeled the SEB Hot Spot at
17◦S and 208◦W (marked "B" on Figures 1, 4, and
5), which had the highest radiance in our entire
dataset. The noise level was measured using the
standard deviation of the number of counts in the
spectral pixel corresponding to a saturated telluric
water line in each order. This was evaluated over
11 spatial pixels over the lowest flux region (the
STZ) and also for 11 spatial pixels off of Jupiter.
Both gave the same result. In Order 16 the sig-
nal to noise ratio (S/N) of the SEB Hot Spot at
2141.6 cm−1 was 2500; the S/N of the zone at
32◦S was 48. In Order 15 where strong H2O lines
occur, the S/N of the same Hot Spot was 1900 at
2012.2 cm−1 ; the S/N of the STZ was 35.
Synthetic spectra were calculated using the
Spectrum Synthesis Program (SSP)
radiative
transfer code as described in Kunde & Maguire
(1974). The input temperature profile was ob-
tained from the Galileo Probe (Seiff et al. 1998).
Line parameters for CH3D and other 5-µm ab-
sorbers are from GEISA 2003 (Jacquinet-Husson et al.
2005).
Parameters for CH3D-H2 and CH3D-
He broadening have been measured in the lab
(Boussin et al. 1999; Lerot et al. 2003; F´ejard et al.
2003). We used a broadening coefficient of
0.0613 cm−1 /atm (296/T)0.5 for CH3D colliding
with a mixture of 86.3% H2 and 13.6% helium, as
measured by the Galileo Probe (von Zahn et al.
1998). Pressure-induced H2 coefficients were ob-
tained using laboratory measurements at 5 µm by
Bachet et al. (1983) and the formalism developed
by Birnbaum & Cohen (1976).
The base of the model was set to 20 bars, 416K
for the SEB Hot Spot to ensure that the base was
well below the level where Jupiter's atmosphere
becomes optically thick due to pressure-induced
H2 opacity. For the zone at 32◦S, we investigated
lower boundaries at 2, 4, 5, and 20 bars to simu-
late opaque NH4SH clouds, H2O clouds, and no
deep clouds in the troposphere. For the Hot Spot,
we calculated a spectrum free of deep clouds for an
emission angle of 18.7◦ and for an H2O mole frac-
tion of 47 ppm for P > 4.5 bars as shown in Fig. 3,
consistent with results from the Galileo Probe
(Wong et al. 2004b). The model was iterated to
fit CH3D (0.18 ppm) and PH3 (0.45 ppm). Our
fit to CH3D in the Hot Spot is very close to the
value (0.16±0.04 ppm) derived by Lellouch et al.
(2001) using spectra at 8.6 µm from the In-
frared Space Observatory. We then convolved
the Jupiter spectrum to 0.02 cm−1 , doppler-
shifted it by 25.2 km/sec, multiplied by the calcu-
5
lated transmittance above Mauna Kea, convolved
it to 0.14 cm−1 to match the NIRSPEC data,
and multiplied the spectrum by 0.332 to simu-
late the transmission of upper cloud layers on
Jupiter to match the observed continuum. As
shown in Fig. 4, the synthetic spectrum fits Re-
gion B, the SEB Hot Spot spectrum without any
deep cloud opacity (P > 2 bars). All 4 CH3D fea-
tures,
including the wing between 2143.2 and
2143.6 cm−1 are matched.
The zone model is more complicated. A ra-
diative transfer model was used to calculate spec-
tra for an emission angle of 33.4◦. Above each
lower boundary, the gas composition was set to
0.18 ppm CH3D. For the zone model with no
deep clouds, the Galileo Probe H2O profile was
used. For models with opaque clouds at 2, 4, and
5 bars, a saturated profile of H2O was used (see
Fig. 3). The mole fraction of PH3 was iterated to
a value of 0.7 ppm to match the absorption fea-
ture at 2143 cm−1 for lower boundaries at 4 and
5 bars. For the zone model without deep clouds,
the PH3 abundance in the STZ was iterated to
a value of 0.45 ppm. Thus, PH3
can also be
used to discriminate between deep cloud models,
as we describe below. A larger PH3 mole frac-
tion (0.7 ppm) above an opaque cloud layer at 4
bars yields the same PH3 absorption as a smaller
mole fraction (0.45 ppm) above the (deeper) level
where the atmosphere becomes opaque due to H2-
H2 opacity. The synthetic spectrum was split into
two parts. The reflected solar component was cal-
culated using the transmittances above a reflect-
ing layer at 300 mbar and for an upper cloud re-
flectance of 9%. It was convolved to 0.02 cm−1 res-
olution, doppler-shifted, and multiplied by the
ATMOS solar spectrum. The thermal component
was convolved, doppler-shifted, and multiplied by
transmittances of 0.393, 0.0269, 0.016, and 0.0038
for models with a base of 2, 4, 5, and 20 bars, re-
spectively, and added to the reflected spectrum.
This,
in turn, was multiplied by the transmis-
sion above Mauna Kea and finally convolved to
0.14 cm−1 resolution.
Due to the numerous parameters in this model,
we illustrate the parameter sensitivity of the
model for the reflected solar and thermal com-
ponents of the zone model separately in Fig. 5.
We treat the reflective upper cloud as spectrally
grey.
Ice components such as NH3, NH4SH,
6
and H2O have been seen at other wavelengths
(e.g., Brooke et al.
(1998), Simon-Miller et al.
(2000), Baines et al. (2002), Wong et al. (2004a),
Sromovsky & Fry (2010)). However, our spec-
tral windows do not include significant ice ab-
sorption features. The imaginary indices of re-
fraction of NH3 (Martonchik et al. 1984) and
NH4SH (Howett et al. 2007) are on the order
of 10−3 at the wavelengths studied here, much
lower than values closer to unity at the ν3 vibra-
tion transitions for these ices (at 3.0 and 3.4 µm,
respectively). Thus, the spectral features in the
reflected solar component are due to CH3D ab-
sorption lines in the upper troposphere of Jupiter
or Fraunhofer lines in the Sun.
In the top panel of Fig. 5 we show 4 Fraunhofer
lines, marked S. The strength of the strongest Jo-
vian feature at 2141.8 cm−1 when compared with
the spectrum of the Sun acquired by the ATMOS
investigation yields the relative fractions of re-
flected sunlight (0.43) and thermal emission (0.57)
for this portion of Jupiter's South Tropical Zone.
The fraction of reflected sunlight combined with
the calibrated continuum level allows us to derive
an upper cloud reflectance of 9%. There are 5
CH3D absorption features in this spectral range.
Ammonia clouds and hazes limit the penetration
of reflected sunlight to the upper troposphere. We
investigated the pressure of the reflecting layer by
calculating spectra at 100, 300, and 600 mbars.
We adopted a value of 300 mbars for the STZ.
Note that in order to match the narrow absorption
cores of CH3D in a model that consists of the sum
of two components, there are a family of solutions
for the pressure level of the solar reflecting layer
and for the thermal cloud top pressure. Increasing
the pressure of the reflecting layer from 300 to 600
mbars may be compensated by, for example, de-
creasing the (deep) cloud top pressure from 4 bars
to 2 bars in the thermal component. However, we
can exclude this possibility by studying the wing
of the CH3D feature between 2143.2 and 2143.6
cm−1. The top panel shows that the reflected solar
spectrum is flat over this range. In contrast, each
thermal model has a different CH3D wing line
slope. We now return to Fig. 4 which compares
the sum of the reflected and thermal components
of each model to the observed STZ spectrum. We
can exclude models with an opaque NH4SH cloud
at 2 bars (blue curve) as well as the model with no
deep cloud (red curve). Note that the calculated
radiances for both of these models lie well outside
the error bars of the zone spectrum. The model
with a deep cloud at 5 bars fits portions of the
spectrum but the model with an opaque cloud at
4 bars (green curve) provides a better overall fit
to the spectrum.
The NIRSPEC spectral bandpass for Order 16
covers more than just this particular CH3D ab-
sorption feature. In Fig. 6 we compare the STZ
spectrum with the same zone models for a spec-
tral region adjacent to the one displayed in Figs.
4 and 5. Here there are three telluric H2O lines.
To the left of each telluric line is a Jovian water
line red-shifted by 0.18 cm−1 due to the relative
velocity of Jupiter with respect to the Earth. The
wings of these water lines permit us to discrim-
inate between models. The best fit requires an
opaque cloud between 4 and 5 bars.
Next, we compare our retrieved PH3
abun-
dances in the 4 to 8-bar level of the STZ with
measurements of PH3 at 1 bar from the Cassini
flyby of Jupiter in January 2001.
Irwin et al.
(2004) retrieved PH3 mole fractions on Jupiter
ranging from 0.9 to 1.5 ppm using zonal aver-
ages of CIRS spectra at 9 µm between 60◦S and
60◦N. The retrieved PH3 mole fraction at 32◦S
was 1.0±0.2 ppm. Phosphine falls off with height
in the upper troposphere due to ultraviolet pho-
tolysis, but its mole fraction is not expected to
change between 1 and 8 bars. There remain dis-
crepancies between PH3
retrievals at 5 and 9
µm (Fletcher et al. 2009). Nevertheless, better
agreement for the abundance of PH3 at 32◦S is
achieved for a model with opaque clouds near 4 to
5 bars (0.7 ppm) than for a zone model with no
deep cloud (0.45 ppm).
Thus, using three independent arguments that
are based on: a) the slope of the CH3D line wings,
b) the slope of the H2O line wings, and c) the
derived mole fractions of PH3, we conclude that
there must be significant cloud opacity between 4
and 5 bars in the STZ at 32◦S. Based on the tem-
peratures at these pressure levels (257K to 275K),
thermochemical models predict that this is a water
cloud.
In Fig. 7 we compare the spectrum of the SEB
Hot Spot marked B at 2016 cm−1 (4.96 µm) with 3
different models calculated using the vertical pro-
files of H2O shown in Fig. 3. The parameters
7
are the same as those used to fit the spectrum at
2144 cm−1 except for H2O and NH3. We initially
used a vertical profile of NH3 derived from ab-
sorption of the radio signal from the Galileo Probe
(Folkner et al. (1998) as modified by Hanley et al.
(2009)). We found that a scaling factor of 2 was
required to fit the NH3 absorption features shown
in Fig. 7. This corresponds to NH3 mole frac-
tions ranging from 240 ppm at 2 bars to 600 ppm
at 4.5 bars. This is significantly larger than val-
ues derived from ground-based microwave observa-
tions (see Sault et al. (2004)). However, we have
not used the much stronger absorption features
in Order 14 (5.3 µm) to constrain the NH3 ver-
tical profile. Thus, the NH3 mole fractions re-
ported here should be regarded as preliminary.
The three different vertical profiles of H2O in
Fig. 3, and used in our calculations, were based on
the mass spectrometer data on the Galileo Probe.
Niemann et al. (1998) reported an upper limit to
H2O of 0.8 ppm at 2.7 bars. Wong et al. (2004b)
reported a mole fraction of H2O of 40±13 ppm at
11.0-11.7 bars on Jupiter. The only adjustable pa-
rameter in our model was the pressure at which the
H2O mole fraction increased from 0.8 to 47 ppm,
as shown in Fig. 3. The best fit is for a pressure
of 4.5 bars. Assuming that the SEB Hot Spot is
similar to the one that Galileo entered, this data
point provides a useful measurement of the depth
at which dynamical processes have dried out Hot
Spots on Jupiter. This depth compares well with
the depth determined by Sault et al. (2004) for the
NH3 abundance in hot spots from microwave ob-
servations.
We next modeled the zone marked A at
2016 cm−1. We used the same procedure as was
used to model the zone spectrum at 2144 cm−1.
We explored the same set of models with cloud
tops at 2, 4, and 5 bars, as well as the model with
no deep cloud. This portion of Jupiter's spectrum
is not sensitive to the deep cloud structure. Using
the results from fitting CH3D, we adopted a model
with an opaque cloud top at 4.0 bars and a satu-
rated H2O profile above. The NH3 mole fraction
was iterated until a rough fit was achieved using
a value of 250 ppm between 0.75 and 4.0 bars and
a saturated value for P < 0.75 bars. Without
any obvious Fraunhofer lines in this spectral re-
gion to constrain the reflected solar component,
we assumed the same upper cloud reflectance of
9% as at 2144 cm−1 and we obtained an upper
cloud transmittance of 0.0542. The transmittance
is therefore less certain than the value derived at
2144 cm−1. The principal conclusion from this
spectral region is that the observed H2O line
profiles in the STZ are slightly broader than the
telluric H2O features, but narrower than H2O fea-
tures in the SEB Hot Spot. A saturated H2O pro-
file for P < 4.0 bars provides a satisfactory fit to
the observed spectrum.
1.0
2.0
)
s
r
a
b
(
e
r
u
s
s
e
r
P
5.0
10.0
Water Vapor Vertical Profiles
Galileo
Upper Limit
Saturated Profile
SEB Hot Spot Profiles
Galileo
Probe
20.0
10-7
10-6
10-5
10-4
10-3
H2O mole fraction
Fig. 3. -- Vertical profiles of H2O used to calculate syn-
thetic spectra in Figures 4 and 5.
8
Sensitivity to Cloud Top Pressure Level
1.4 10-7
1.2 10-7
)
1
-
m
c
/
1
-
r
s
2
-
m
c
t
t
a
W
(
e
c
n
a
d
a
R
i
t
o
p
S
t
o
H
6.0 10-8
4.0 10-8
2.0 10-8
1.0 10-7
8.0 10-8
0.0
S
PH3
Zone 32oS
A
Cloud 2 bars
Cloud 4 bars
Cloud 5 bars
No deep cloud
B
Hot Spot 17oS
No deep cloud
Mauna Kea Trans
2142.0
2143.0
H2O
T
2145.0
2146.0
CH3D
2144.0
Wavenumber (cm-1)
2.5 10-9
2.0 10-9
1.5 10-9
1.0 10-9
5.0 10-10
Z
o
n
e
0.0
Fig. 4. -- The pressure level of deep clouds on Jupiter as constrained by CH3D line profiles observed in Order 16. An SEB
Hot Spot (gray curve) at 17◦S (labeled B in Fig. 1) is fitted without any clouds at pressures greater than 2 bars (pink curve).
Hot Spot error bars would be smaller than the thickness of the gray curve. The spectrum of the STZ at 32◦S (points with
error bars, and labeled A in Fig. 1) was modeled using opaque clouds at 2, 4, 5 bars and a model without any deep clouds. We
may exclude an opaque NH4SH cloud at 2 bars due to the poor fit of the blue curve to the observed spectrum. The best fit
requires an opaque cloud between 4 and 5 bars. Thermochemical models predict that cloud opacity at this pressure level is due
to a water cloud. Note the factor of 50 difference in radiance scales between the Hot Spot and the zone spectra. The calculated
transmittance above Mauna Kea for 2 mm precip H2O is shown as a dashed blue line. T denotes a telluric H2O line and S
denotes a solar (Fraunhofer) line.
9
Reflected Solar Component of South Tropical Zone
PH3
S
S
S
S
CH3D CH3D
CH3D
CH3D
CH3D
Zone 32oS
A
Reflect 100 mbar
Reflect 300 mbar
Reflect 600 mbar
T
H2O
2142.0
2143.0
2144.0
2145.0
2146.0
Wavenumber (cm-1)
Thermal Component of South Tropical Zone
PH3
CH3D
CH3D
CH3D
CH3D
T
H2O
Zone 32oS
A
Cloud Top 2 bars
Cloud Top 4 bars
Cloud Top 5 bars
No Deep Cloud
1.0 10-9
5.0 10-10
0.0
R
e
f
l
e
c
t
e
d
R
a
d
i
a
n
c
e
1.5 10-9
1.0 10-9
5.0 10-10
0.0
T
h
e
r
m
a
l
i
R
a
d
a
n
c
e
i
e
c
n
a
d
a
R
e
n
o
Z
l
a
c
i
p
o
r
T
h
t
u
o
S
i
e
c
n
a
d
a
R
e
n
o
Z
l
i
a
c
p
o
r
T
h
t
u
o
S
2.5 10-9
2.0 10-9
1.5 10-9
1.0 10-9
5.0 10-10
0.0
2.5 10-9
2.0 10-9
1.5 10-9
1.0 10-9
5.0 10-10
0.0
2142.0
2143.0
2144.0
2145.0
2146.0
Wavenumber (cm-1)
Fig. 5. -- Top: The reflected solar component of a radiative transfer model of the STZ at 32◦S. Four Fraunhofer lines are
marked as S. A telluric H2O line is marked as T. A reflecting layer is placed at 100, 300, and 600 mbars in Jupiter's upper
troposphere. One PH3 and five CH3D lines are shown. The narrow cores of the CH3D lines are sensitive to the solar
component. Bottom: The thermal component is shown for models with cloud tops at 2, 4, and 5 bars, and for a model with
no deep cloud. The wing of CH3D between 2143.2 and 2143.6 cm−1 and the PH3 line at 2143.0 cm−1 are due to the thermal
component. Radiance scales are offset to match the continuum of the observed spectrum.
10
2.5 10-9
2.0 10-9
1.5 10-9
1.0 10-9
5.0 10-10
0.0
)
1
-
m
c
/
1
-
r
s
2
-
m
c
t
t
a
W
(
e
c
n
a
d
a
R
i
Sensitivity to Cloud Top Pressure Level
S
H2O
H2O
T
H2O
H2O
T
H2O
A
Zone 32oS
Cloud Top 2 bars
Cloud Top 4 bars
Cloud Top 5 bars
No Deep Cloud
T
H2O
2137.0
2138.0
2139.0
2140.0
2141.0
2142.0
Wavenumber (cm-1)
Fig. 6. -- The spectrum of the STZ at 32◦S and radiative transfer models with opaque clouds at 2, 4, and 5 bars, and a model
without any deep clouds are shown for a spectral region adjacent to the CH3D features. T denotes telluric H2O lines and
S denotes a solar (Fraunhofer) line. Note the 3 Jovian H2O lines red-shifted by 0.18 cm−1 from their telluric counterparts.
The wings of these water lines permit us to discriminate between models. The best fit requires an opaque cloud between 4 and
5 bars.
11
Sensitivity to Water Vapor Vertical Profile
A
Zone 32oS
H2O Sat P<4.0 bars
1.6 10-7
)
1
-
m
c
/
1
-
r
s
2
-
m
c
t
t
a
W
1.2 10-7
8.0 10-8
(
e
c
n
a
d
a
R
i
4.0 10-8
t
o
p
S
t
o
H
0.0
NH3
H2O
H2O
H2O
B
NH3
Hot Spot 17oS
H2O Gal P>3.5 bars
H2O Gal P>4.0 bars
H2O Gal P>4.5 bars
Mauna Kea Trans
2012.0
2014.0
2016.0
2018.0
2020.0
Wavenumber (cm-1)
3.5 10-9
3.0 10-9
2.5 10-9
PH3
2.0 10-9
1.5 10-9
1.0 10-9
5.0 10-10
0.0
Z
o
n
e
R
a
d
Fig. 7. -- Vertical profiles of water vapor mixing ratio as constrained by H2O line profiles observed in Order 15. An SEB Hot
Spot (Region B, gray curve) at 17◦S is fitted without any deep clouds. Hot Spot error bars would be smaller than the thickness
of the gray curve. Three vertical profiles of H2O were calculated for the Hot Spot. The best fit used the Galileo Probe value
for H2O for pressures greater than 4.5 bars (see also Fig. 3). The spectrum of the STZ at 32◦S (Region A, points with error
bars) was fitted using an opaque cloud at 4 bars and a saturated H2O profile above it (purple curve). The H2O lines in the
zone are slightly broader than the telluric features, but much narrower than in the Hot Spot.
12
Latitudinal Variation
4.66 µm Flux
CH3D Eq Width
Fraction Reflected Solar
Possible Water Clouds
Possible Water Clouds
Possible Water Clouds
No Opaque Water Clouds
Thick Upper Clouds
No Opaque Water Clouds
NEB Hot
Confirmed No Opaque Water Clouds
SEB Hot
B
A
Confirmed Water Cloud
Thick Upper Clouds
Possible Water Clouds
Possible Water Clouds
No Opaque Water Clouds
60.0
40.0
20.0
0.0
-20.0
-40.0
-60.0
e
d
u
t
i
t
a
L
c
i
h
p
a
r
g
o
t
e
n
a
l
P
0.0
0.2
0.4
Normalized Quantity
0.6
0.8
1.0
1.2
Fig. 8. -- The flux at 4.66 µm (red curve), the equivalent width of CH3D (black curve), and the the ratio of the equivalent
width of the Fraunhofer line at 2141.8 cm−1 on Jupiter to that in the Sun is shown for each latitude along the central meridian
on Jupiter. Hot Spots are present at 8.5◦N and 17◦S. Latitudes exhibiting minima in CH3D equivalent widths are candidates
for water clouds. Radiative transfer models were used to match a zone spectrum at 32◦S (marked "A") and a Hot Spot spectrum
at 17◦S (marked "B"). The gold curve indicates the fraction due to reflected sunlight. Large values imply thick upper clouds
that attenuate thermal radiation from below and reflect sunlight above. The bar on the right denotes haze reflectivity variation
with latitude obtained from 2.12 µm SCAM images, roughly similar to belt-zone structure seen at visible wavelengths.
13
4. DISCUSSION
This paper describes a technique to determine
water cloud heights using spectrally-resolved line
shapes of CH3D. We have demonstrated with a set
of spectra from characteristic spatial regions that
variation in the water cloud height manifests as a
variation in the equivalent width of CH3D. This
technique can be applied to all regions covered by
the spectrometer slit. We show the CH3D equiv-
alent width as a function of latitude (black curve)
in Fig. 8, highlighting the latitude regions which
may have water clouds. The observations are ex-
tremely valuable both for understanding Jupiter's
atmospheric circulation, as well as its bulk water
abundance (a constraint on planetary formation
scenarios).
However, the equivalent width of CH3D alone
In order to map H2O clouds
is not sufficient.
on Jupiter, we need to measure the latitudinal
profile of both CH3D line shapes as well as the
strength of the Fraunhofer line at 2141.8 cm−1.
In Fig. 8 we plot three quantities as a function
of latitude on Jupiter. First is the radiance inte-
grated over the 4.6-µm bandpass and normalized
to 1.0 as a function of spatial pixel and converted
to planetographic latitude. This is shown in red.
The next quantity is the equivalent width of our
set of CH3D absorption lines integrated between
2143.21 and 2145.04 cm−1(denoted by the hori-
zontal bar in Fig. 2). This is shown in black. The
last quantity plotted in Fig. 8 is the fraction of
reflected sunlight as a function of latitude. This
was obtained by measuring the equivalent width
of the Fraunhofer line at 2141.8 cm−1 as a func-
tion of latitude and dividing it by the value in the
Sun as measured by ATMOS. At latitudes such as
23◦N where this value is close to 1.0, there is in-
sufficient thermal flux to constrain the deep cloud
structure. However, by using adjacent latitudes
where CH3D equivalent widths are also small, and
where the fraction of reflected solar flux < 0.6, we
can infer the presence of water clouds. Thus, we
have indicated "possible water clouds" for the en-
tire region where values of CH3D equivalent width
are small. The vertical bar on the right denotes
the brightness of Jupiter in reflected sunlight at
2.12 µm obtained from SCAM images navigated
onto a planetographic latitude grid. This wave-
length sounds hazes in the upper troposphere and
illustrates Jupiter's belt-zone structure.
The equivalent width of CH3D is sensitive to
two parameters: the fraction of reflected sunlight
and the pressure level of the deep cloud. Consider
two regions on Jupiter. They both lack an opaque
water cloud, but one has thin upper clouds so that
the 5-µm spectrum is 100% thermal. The other
has a thick, reflective upper cloud that attenuates
the thermal component yielding a spectrum that
is 50% thermal and 50% reflected solar.
In the
latter case, the contribution of the reflected solar
continuum will significantly reduce the equivalent
width of the CH3D absorption features that are
present in the thermal component. A low value
of CH3D equivalent width might be interpreted
as due to a deep water cloud, whereas in this case
the small equivalent width was due solely to the
presence of the solar component from an upper
reflective layer. Thus, we have to be cautious in
using the equivalent width of CH3D alone to in-
fer deep cloud structure. Nevertheless, the data
plotted in Figure 8 show interesting structure at
latitudes where the fraction of reflected solar radi-
ation is essentially zero. High latitudes, southward
of 55◦S and northward of 35◦N show interesting
variations in CH3D equivalent widths that can-
not be attributed to reflected sunlight and thus
are related to deep cloud structure.
With the caveat that radiative transfer mod-
eling is required to verify the presence of wa-
ter clouds in regions with small CH3D equiva-
lent widths and significant fractions of reflected
sunlight, we can make the following conclusions:
Where thick water clouds are present at pres-
sures in the 3-5 bar range, the CH3D equiva-
lent width is small, and the black curve (Fig. 8)
tends toward lower values. Regions with thick,
high-altitude (P ≤ 4 bars) water clouds can be
associated with widespread upwelling, since down-
welling would provide the opposite effect of clear-
ing cloud opacity via sublimation. We find the
smallest CH3D equivalent widths in the three low-
latitude zones: the equatorial zone and the north
and south tropical zones. This implies a large ver-
tical extent to the upwelling responsible for form-
ing the thick white clouds observed in zones at
optical wavelengths.
Thick water clouds in zones specifically con-
tradict models of inverted two-layer circulation
within the tropospheric cloud decks. Ingersoll et al.
14
(2000) and Showman & de Pater (2005) suggested
inverted two-layer circulation schemes to explain
widespread ammonia gas depletion even in zones.
In this type of circulation model, regions of up-
welling (thick clouds) in the visible upper levels
(NH3+ NH4SH cloud decks) corresponds to down-
welling at the deeper, hidden water cloud level.
Our results suggest that mass flux into the upper
cloud layers of zones is not dominated by hori-
zontal transport, as in the Showman & de Pater
scenario, but is driven by vertical transport from
below.
Jupiter's circulation in zones therefore
maintains the same sign of upwelling/downwelling
across the full 0.5-5 bar weather layer. The same
sign of upwelling/downwelling over this large an
extent was also derived by de Pater et al. (2010)
from 5-micron bright rings around vortices. These
authors suggested that vortices must extend ver-
tically from at least the 4-7 bar level up to the
tropopause.
Large CH3D equivalent width requires low wa-
ter cloud opacity. Subsidence of dry upper-level
air provides the simplest mechanism. Figure 8
shows that at low latitudes, the clearest deep at-
mosphere (high values of the black curve) occur
where the 5-µm flux is highest (red curve). The
very highest flux levels correspond to an atmo-
sphere largely devoid of clouds of all three types,
H2O, NH3, and NH4SH, again suggesting down-
welling circulation that spans the full 0.5-5 bar
range at least. However, the width (in latitude)
of the 5-µm flux peak is very narrow, compared to
the width of the NEB/SEB regions with low water
cloud opacity. Thus, over a broad latitude range
with little or no water cloud opacity, the atmo-
sphere is characterized by both cloudy and cloud-
free conditions at the upper levels. Perhaps a two-
layer circulation model with a vertical flow reversal
layer between the water and NH3/ NH4SH clouds
can be relevant on these regional scales, if not for
the entire planet.
The rapid "fading" or whitening of the South
Equatorial Belt between 2008 and 2010 was in-
terpreted by Fletcher et al. (2011) as due to en-
hanced upwelling of ammonia-rich air followed by
condensation. Using CH3D line profiles and ab-
sorption lines of NH3 and H2O in the 5-µm win-
dow, we now have the capability of measuring
cloud structure and volatile abundances in the
NEB and SEB at the 4 to 8 bar level. Any fu-
ture changes in the appearance of the belts can
now be investigated over the full 0.5 to 8-bar range
of Jupiter's troposphere using spectroscopy at 4.6
and 8.6 µm. By studying changes in each cloud
layer separately we will have a much better un-
derstanding of the dynamics below Jupiter's visi-
ble clouds.
The high latitudes also show high CH3D equiv-
alent widths. Here, more detailed modeling will
be needed to disentangle geometric effects caused
by viewing geometry, as well as changes in at-
mospheric scale height in a rapidly-rotating non-
spherical planet. But the data do suggest a differ-
ent paradigm in deep cloud structure at high lat-
itudes (polewards of 40◦). We note, though, that
5-micron images at high spatial resolution show a
lot of structure at these latitudes (de Pater et al.
2011), and microwave images show an overall low
NH3 abundance (de Pater 1986). A combination
of 5-µm imaging, microwave imaging, and 5-
µm spectroscopy of Jupiter's polar regions would
be extremely useful to investigate these interesting
cloud features.
We now return to the interpretation of 5-
µm spectra of Jupiter acquired by the KAO,
Voyager/IRIS, and Galileo/NIMS. It now seems
clear that the airborne observations of Jupiter
were flux-weighted by Hot Spots. The abun-
dances of H2O and other molecules derived by
Bjoraker et al. (1986a,b) pertain to Hot Spots,
but not to Jupiter as a whole. The lower bound-
ary for Hot Spots is in fact due to pressure-induced
H2 opacity, rather than an opaque water cloud,
as proposed by Carlson et al. (1992). However,
the model proposed by Carlson et al. does ap-
pear to apply to Jupiter's zones, at least in the
regions marked in Fig. 8 as candidates for wa-
ter clouds. Similarly, models of Galileo/NIMS
spectra of the Equatorial Zone by Drossart et al.
(1998) are probably accurate while models of Hot
Spots that included water clouds (e.g. Nixon et al.
(2001)) will need to be revised.
This technique provides a constraint on Jupiter's
deep water abundance and therefore its O/H ra-
tio. The base of the water cloud is sensitive to the
abundance of water because higher abundances
lead to condensation at deeper levels. The data
provide the level of the cloud top, not the cloud
base. Since the top is at higher altitude than the
base, cloud top constraints provide lower limits to
15
the pressure of the cloud base, or, lower limits to
the deep abundance of water. A spectrum requir-
ing a water cloud at P ≥ 5 bar would establish a
supersolar enrichment of water in Jupiter, better
constraining planetary formation models (Wong
et al. 2008).
Figure 4 shows the effect of water cloud pres-
sure level on model fits to a cloudy zone spec-
trum at 32◦S. The spectrum is best fit by a water
cloud with a top between 4-5 bars. The cloud
base is therefore found at P > 4 − 5 bar. Fol-
lowing Fig. 1 in Wong et al. (2008), a cloud base
at 4-5 bar corresponds to O/H ratios 0.33-1.1×
solar (corrected to the new solar O/H ratio of As-
plund et al. 2009). Our observations thus pro-
vide a lower limit to Jupiter's water abundance
of 0.33-1.1× solar. This result is consistent with
the Galileo Probe lower limit of 0.48±0.16× solar
(Wong et al. 2004), and therefore does not provide
any new constraint on the water abundance. In fu-
ture work, we will search for spectra that require
an even deeper water cloud.
Knowledge of the deep cloud structure permits
us to retrieve abundances of H2O, NH3, and other
5-µm absorbers more accurately. This will enable
studies of dynamics below Jupiter's visible cloud
layers. Additional observations will permit us to
observe discrete cloud features such as the Great
Red Spot and Oval BA.
The data presented were obtained at the W.
M. Keck Observatory, which is operated as a sci-
entific partnership among the California Institute
of Technology, the University of California, and
the National Aeronautics and Space Administra-
tion. The Observatory was made possible by the
generous financial support of the W. M. Keck
Foundation. The authors extend special thanks
to those of Hawaiian ancestry on whose sacred
mountain we are privileged to be guests. Without
their generous hospitality, none of the observa-
tions presented would have been possible. We
also would like to thank Linda Brown for steering
us to the latest broadening coefficients for CH3D.
This research was supported by the NASA Plan-
etary Astronomy (PAST) Program grant number
NNX11AJ47G, NNX14AJ43G, NNX15AJ41G,
and NASA Outer Planets Research Program grant
number NNX11AM55G.
Facilities: Keck Observatory, Voyager (IRIS).
16
REFERENCES
Adriani, A., Filacchione, G., Di Iorio, T., et al.
2014, Space Sci. Rev., doi:10.1007/s11214-014-
0094-y
Asplund, M., Grevesse, N., Sauval, A. J., & Scott,
P. 2009, ARA&A, 47, 481
Bachet, G., Cohen, E. R., Dore, P., & Birnbaum,
G. 1983, Canadian Journal of Physics, 61, 591
Baines, K. H., Carlson, R. W., & Kamp, L. W.
2002, Icarus, 159, 74
Birnbaum, G., & Cohen, E. R. 1976, Canadian
Journal of Physics, 54, 593
Bjoraker, G. L., Hewagama, T., & Orton, G. S.
2002, in Bulletin of the American Astronomical
Society, Vol. 34, AAS/Division for Planetary
Sciences Meeting Abstracts #34, 874
Bjoraker, G. L., Larson, H. P., & Kunde, V. G.
1986a, ApJ, 311, 1058
-- . 1986b, Icarus, 66, 579
Boussin, C., Lutz, B. L., Hamdounia, A., & de
Bergh, C. 1999, J. Quant. Spec. Radiat. Transf.,
63, 49
Brooke, T. Y., Knacke, R. F., Encrenaz, T., et al.
1998, Icarus, 136, 1
Carlson, B. E., Lacis, A. A., & Rossow, W. B.
1992, ApJ, 388, 648
de Pater, I. 1986, Icarus, 68, 344
de Pater, I., DeBoer, D., Marley, M., Freedman,
R., & Young, R. 2005, Icarus, 173, 425
de Pater, I., Wong, M. H., de Kleer, K., et al.
2011, Icarus, 213, 559
de Pater, I., Wong, M. H., Marcus, P., et al. 2010,
Icarus, 210, 742
Drossart, P., Roos-Serote, M., Encrenaz, T., et al.
1998, J. Geophys. Res., 103, 23043
Farmer, C. B. 1994, in IAU Symposium, Vol. 154,
Infrared Solar Physics, ed. D. M. Rabin, J. T.
Jefferies, & C. Lindsey, 511
Farmer, C. B., & Norton, R. H. 1989, A high-
resolution atlas of the infrared spectrum of the
sun and the earth atmosphere from space. A
compilation of ATMOS spectra of the region
from 650 to 4800 cm−1 (2.3 to 16 µm). Vol. I.
The sun.
F´ejard, L., Gabard, T., & Champion, J.-P. 2003,
Journal of Molecular Spectroscopy, 219, 88
Fletcher, L. N., Orton, G. S., Teanby, N. A., &
Irwin, P. G. J. 2009, Icarus, 202, 543
Lerot, C., Walrand, J., Blanquet, G., Bouanich,
J.-P., & Lep`ere, M. 2003, Journal of Molecular
Spectroscopy, 219, 329
Martonchik, J. V., Orton, G. S., & Appleby, J. F.
1984, Appl. Opt., 23, 541
McLean, I. S., Becklin, E. E., Bendiksen, O., et al.
1998, in Society of Photo-Optical Instrumenta-
tion Engineers (SPIE) Conference Series, Vol.
3354, Infrared Astronomical Instrumentation,
ed. A. M. Fowler, 566 -- 578
Fletcher, L. N., Orton, G. S., Rogers, J. H., et al.
Niemann, H. B., Harpold, D. N., Atreya, S. K.,
2011, Icarus, 213, 564
et al. 1992, Space Sci. Rev., 60, 111
Folkner, W. M., Woo, R., & Nandi, S. 1998,
Niemann, H. B., Atreya, S. K., Carignan, G. R.,
J. Geophys. Res., 103, 22847
et al. 1998, J. Geophys. Res., 103, 22831
Friedson, A. J. 2005, Icarus, 177, 1
Hanley, T. R., Steffes, P. G., & Karpowicz, B. M.
2009, Icarus, 202, 316
Hersant, F., Gautier, D., & Lunine, J. I. 2004,
Planet. Space Sci., 52, 623
Howett, C. J. A., Carlson, R. W., Irwin, P. G. J.,
& Calcutt, S. B. 2007, Journal of the Optical
Society of America B Optical Physics, 24, 126
Ingersoll, A. P., Gierasch, P. J., Banfield, D.,
Vasavada, A. R., & Galileo Imaging Team.
2000, Nature, 403, 630
Irwin, P. G. J., Parrish, P., Fouchet, T., et al.
Nikitin, A., Champion, J. P., Tyuterev, V. G., &
Brown, L. R. 1997, Journal of Molecular Spec-
troscopy, 184, 120
Nixon, C. A., Irwin, P. G. J., Calcutt, S. B., Tay-
lor, F. W., & Carlson, R. W. 2001, Icarus, 150,
48
Orton, G., Ortiz, J. L., Baines, K., et al. 1996,
Science, 272, 839
Ragent, B., Colburn, D. S., Rages, K. A., et al.
1998, J. Geophys. Res., 103, 22891
Roos-Serote, M., Atreya, S. K., Wong, M. K., &
Drossart, P. 2004, Planet. Space Sci., 52, 397
2004, Icarus, 172, 37
Sault, R. J., Engel, C., & de Pater, I. 2004, Icarus,
Jacquinet-Husson, N., Scott, N. A., Chedin, A.,
et al. 2005, Journal of Quantitative Spec-
troscopy and Radiative Transfer, 95, 429
168, 336
Showman, A. P., & de Pater, I. 2005, Icarus, 174,
192
Janssen, M. A., Hofstadter, M. D., Gulkis, S.,
Simon-Miller, A. A., Conrath, B., Gierasch, P. J.,
et al. 2005, Icarus, 173, 447
& Beebe, R. F. 2000, Icarus, 145, 454
Kunde, V. R., & Maguire, W. C. 1974, Jour-
nal of Quantitative Spectroscopy and Radiative
Transfer, 14, 803
Sromovsky, L. A., & Fry, P. M. 2010, Icarus, 210,
230
Weidenschilling, S. J., & Lewis, J. S. 1973, Icarus,
Larson, H. P., Fink, U., Treffers, R., & Gautier,
20, 465
III, T. N. 1975, ApJ, 197, L137
Lellouch, E., B´ezard, B., Fouchet, T., et al. 2001,
A&A, 370, 610
Wong, M. H., Atreya, S. K., Kuhn, W. R., Ro-
mani, P. N., & Mihalka, K. M. 2015, Icarus,
245, 273
17
Wong, M. H., Bjoraker, G. L., Smith, M. D.,
Flasar, F. M., & Nixon, C. A. 2004a,
Planet. Space Sci., 52, 385
Wong, M. H., Lunine, J., Atreya, S. K., et al.
2008, in Oxygen in the Solar System, Reviews
in Mineralogy and Geochemistry, ed. MacPher-
son, G. J., Mittlefehldt, D. W., Jones, J. & Si-
mon, S. B., Vol. 68, 219 -- 246
Wong, M. H., Mahaffy, P. R., Atreya, S. K., Nie-
mann, H. B., & Owen, T. C. 2004b, Icarus, 171,
153
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
18
|
1910.04720 | 1 | 1910 | 2019-10-10T17:24:52 | Doubly Synchronous Binary Asteroid Mass Parameter Observability | [
"astro-ph.EP"
] | The full two-body problem (F2BP) is often used to model binary asteroid systems, representing the bodies as two finite mass distributions whose dynamics are influenced by their mutual gravity potential. The emergent behavior of the F2BP is highly coupled translational and rotational mutual motion of the mass distributions. A large fraction of characterized binary asteroids appear to be at, or near, the doubly synchronous equilibrium, which occurs when both bodies are tidally-locked and in a circular co-orbit. Stable oscillations about this equilibrium can be shown, for the nonplanar system, to be combinations of seven fundamental frequencies of the system and the mutual orbit rate. The fundamental frequencies arise as the linear periods of center manifolds identified about the equilibrium which are heavily influenced by each body's mass parameters. We leverage these eight dynamical constraints to investigate the observability of binary asteroid mass parameters via dynamical observations. This is accomplished by deriving a relationship between the fundamental frequencies and mass parameters for doubly synchronous systems. This relationship allows us to show the sensitivity of the dynamics to changes in the mass parameters, first for the planar dynamics, and then for the nonplanar dynamics. In so doing we are able to predict the idealized estimation covariance of the mass parameters based on observation quality and define idealized observation accuracies for desired mass parameter certainties. We apply these tools to 617 Patroclus, a doubly synchronous Trojan binary and flyby target of the LUCY mission, as well as the Pluto and Charon system in order to predict mutual behavior | astro-ph.EP | astro-ph |
Doubly Synchronous Binary Asteroid Mass Parameter
Observability$
Alex B. Davisa,∗, Daniel J. Scheeresa
aDepartment of Aerospace Engineering Science, University of Colorado Boulder, CO
80309, United States
Abstract
The full two-body problem (F2BP) is often used to model binary asteroid sys-
tems, representing the bodies as two finite mass distributions whose dynamics
are influenced by their mutual gravity potential. The emergent behavior of
the F2BP is highly coupled translational and rotational mutual motion of
the mass distributions. A large fraction of characterized binary asteroids
appear to be at, or near, the doubly synchronous equilibrium, which occurs
when both bodies are tidally-locked and in a circular co-orbit. Stable oscil-
lations about this equilibrium can be shown, for the nonplanar system, to be
combinations of seven fundamental frequencies of the system and the mutual
orbit rate. The fundamental frequencies arise as the linear periods of cen-
ter manifolds identified about the equilibrium which are heavily influenced
by each body's mass parameters. We leverage these eight dynamical con-
straints to investigate the observability of binary asteroid mass parameters
via dynamical observations. This is accomplished by deriving a relationship
between the fundamental frequencies and mass parameters for doubly syn-
chronous systems. This relationship allows us to show the sensitivity of the
dynamics to changes in the mass parameters, first for the planar dynamics,
and then for the nonplanar dynamics.
In so doing we are able to predict
the idealized estimation covariance of the mass parameters based on obser-
$This material is based upon work supported by the National Science Foundation Grad-
uate Research Fellowship Program under Grant No. DGE 1650115. Any opinions, findings,
and conclusions or recommendations expressed in this material are those of the author(s)
and do not necessarily reflect the views of the National Science Foundation.
∗Corresponding author
Email addresses: [email protected] (Alex B. Davis),
[email protected] (Daniel J. Scheeres)
Preprint submitted to Icarus
October 11, 2019
vation quality and define idealized observation accuracies for desired mass
parameter certainties. We apply these tools to 617 Patroclus, a doubly syn-
chronous Trojan binary and flyby target of the LUCY mission, as well as
the Pluto and Charon system in order to predict mutual behaviors of these
doubly synchronous systems and to provide observational requirements for
these systems mass parameters.
Keywords: Asteroids, dynamics, Asteroids, rotation, Satellites of asteroids,
1. Introduction
In the past three decades binary asteroids have been discovered through-
out the solar system and are believed to make up a significant portion of
many small body populations. Observers have identified 48 transneptunian
binaries and place binaries at approximately 16% of near-Earth asteroids
(NEAs); the majority of equal mass binaries in these populations are ex-
plected to be doubly synchronous[1][2][3]. Radar observations of Near Earth
Asteroids (NEAs), such as 1999 KW4 and 65803 Didymos, have generated
shape models for a number of these bodies. Such information provides a
basis for the application of the full two-body problem (F2BP) to study the
behavior of these systems[4][5]. The F2BP describes the dynamical inter-
actions of two mass distributions, which result in coupled translational and
rotational motion due to the mutual gravity potential between asymmet-
ric mass distributions. The dynamics, stability and the effects of mass and
spin for the F2BP have been studied extensively by Maciejewski, Tricario,
Scheeres, and Boue and Laskar amongst others[6][7][8][9]. Work by Fahne-
stock and Scheeres implemented a polyhedral formulation of the F2BP dy-
namics to simulate the behavior of 1999 KW4 based on a series of radar
observations[4]. Later work by Naidu adapted the inertia integral implemen-
tation of the F2BP developed by Ashenberg to study the spin behavior of
a sampling of 10 well-observed binary NEAs[5][10]. Both studies provided
valuable insight into the dynamical behavior of binaries, however they were
limited by the computational burden or limited expansion order inherent to
their implementation of the F2BP. Recently Hou et al. derived a recursive
approach to the F2BP which enables much more computationally efficient
simulation of the F2BP[11]. The improvements in computationally efficiency
also open the door to study mass distribution sensitivity of binary system
dynamics.
2
As observers continue to study binary systems in more detail it is impor-
tant to understand how their mass parameters may influence the observed
dynamics and affect assumptions made about the system behavior. We ana-
lyze these effects by applying estimation techniques based on idealized obser-
vations of the translational and rotational coupling inherent in a doubly syn-
chronous binary system. Such an approach, while unprecedented for binary
asteroids, has been leveraged during missions to small bodies and asteroid
flybys of the Earth. Most notably, Takahashi and Scheeres used observations
of the spin state of 4179 Toutatis during several Earth flybys to estimate
its moments of inertia[12]. At Vesta the Dawn mission was able to use the
spacecraft's orbital behavior to place constraints on the interior structure
and mass distribution of Vesta and similar plans have been made for the
OSIRIS-Rex mission to Bennu[13][14]. Development of these capabilities for
binaries would enable more reliable measurements and more robust mission
planning for upcoming missions to binary asteroids, such as the LUCY mis-
sion which will fly by the 617 Patroclus binary system and the AIDA mission
which will impact and observe 65803 Didymos.
In this paper we leverage a new formulation of the F2BP dynamics devel-
oped in Hou et al. to explore the ideal observability of binary asteroids mass
parameters from observations of the dynamics. The Hou formulation of the
F2BP is selected because its recursive form provides improved computation
speed and ease of manipulation[11]. To understand the sensitivity of the
binary systems to mass parameters we study the observability of each mass
parameter. To maintain simplicity of the investigation the simulated binary
system is assumed to be in or oscillating near a doubly synchronous equilib-
rium, meaning both bodies are near tidally locked in a relaxed orbit about
their mutual center of mass. We also assume knowledge of the center of mass
and principal axes of the target bodies so as to better study the direct effects
of the mass parameters themselves on the observable dynamics. Thus, this
paper provides ideal limits on the observability of the mass parameters based
purely on observations of a binary system's mutual motion.
Our analysis is primarily applied to the 617 Patroclus system, a Trojan
asteroid believed to be a near equal mass doubly synchronous binary system
made up of two nearly ellipsoidal bodies. 617 Patroclus thus represents a
realistic manifestation of the doubly synchronous assumptions made in this
dynamical analysis. Additionally as a flyby target for the LUCY mission,
this binary has been the target of a number of recent observation campaigns.
The physical parameters describing 617 Patroclus implemented for this study
3
Table 1: Physical Properties of The Doubly Synchronous 617 Patroclus modeled in this
study
Density
ρ [kg/m−3]
881
Primary
Secondary
Orbit
a [km] b [km]
63.5
58.5
c [km]
49
a [km] b [km]
58.5
54
c [km]
45
r [km] Period [days]
664.6
4.41
are listed in Table 1 and based on the results of the 2013 stellar occultation
study performed by Buie et al. In line with the analysis of Buie et al. we as-
sume the system to be homogenous of constant density[15]. Additionally, we
apply this technique to the Pluto-Charon system, exploring its applicability
to near-spherical systems.
In this paper, we will first rigorously define the F2BP, the mutual gravity
potential and the doubly synchronous equilibrium. With this information
we explore the dynamics of the planar form of the F2BP; analyzing the fun-
damental frequencies of the planar dynamics and their observability. This
analysis is then expanded to the nonplanar form of the dynamics and an ide-
alized mass parameter estimation method is developed and explored. Finally,
we examine the case of near-spherical binaries by applying our methods to
the Pluto-Charon system.
Throughout the paper we refer to observations of a binary system's fun-
damental frequencies, which we assume come from an idealized fictitious
observer. The fictitious observer is imagined to receive regular resolved im-
ages of the system of interest, such that the dynamics associated with each
fundamental frequency can be explicitly measured. The reasons for this ap-
proach are twofold. Firstly, this approach allows us to perform a best-case
analysis and provide an upper bound of what is achievable with the described
methodology. Secondly, it allows us to avoid systemic and stochastic errors
from a given measurement type, such as the effect of uncertainty in shape
and surface properties on light curve data, which we view to be outside the
scope of this work.
2. The Full Two-Body Problem
In order to model the dynamics of binary asteroids we implement a simu-
lation of the F2BP. The F2BP describes the mutual gravitational interactions
between two arbitrary mass distributions, as illustrated in Fig. 1. To model
the mutual gravity of the mass distributions the influence of each infinitesi-
4
mal mass element of both bodies on all other mass elements in the opposing
body must be accounted for. The self gravitational potential is ignored in
this implementation of the F2BP as both bodies are assumed to be rigid for
the time scales of interest and thus their self potentials are constant. The
generic form of the mutual gravity potential is a double volume integral over
both mass distributions, where d is the the distance between an finite mass
element of the primary, dmA, and the secondary, dmB.
1
d
dmAdmB
(1)
A
B
(cid:90)
(cid:90)
U = −G
Figure 1: Diagram of initial system geometry. x represents the principal frame of the
primary, A, x(cid:48) is the principal frame of the secondary, B, and n is the inertial frame.
The finite mass elements dmA and dmB must be integrated over the full body using their
locations relative to the body centers of mass, (cid:126)ρ and (cid:126)ρ(cid:48), to compute the mass parameters
and mutual gravity potential.
In the Hou et al. reformulation the mutual gravity potential is expressed
as a function of the relative separation magnitude, R, and the attitude of
the two approximately modeled asteroids, C A and C B; where the attitude
matrices represent a mapping from the indicated body's principal frame into
the inertial frame. At the second degree and order of the mass distribution,
the mutual potential is:
5
𝒙"𝒙"′𝒗𝑨𝑩𝝎𝝎′𝒏"𝟏𝒏"𝟐𝒏"𝟑𝒓𝝆𝝆′𝒅𝒎1𝒅𝒎2𝒅(cid:16)
− G
2R3
U = −GMAMB
2R5 (cid:126)r •(cid:16)
3G
R
+
MBTr(I A) + MATr(I B)
(2)
MBC AI AC T
A + MAC BI BC T
B
The second order mass distribution terms are the principally aligned in-
ertia tensors, I A and I B, expressed as linear combinations of the inertia
integrals:
(cid:17)
(cid:17) • (cid:126)r
T 0,2,0
i
I i = Mi
i
+ T 0,0,2
0
0
T 2,0,0
i
0
+ T 0,0,2
0
i
0
T 2,0,0
i
+ T 0,2,0
i
(3)
The inertia integrals capture the mass distributions up to a truncation
order N, here selected as the second order. In their generic form the inertia
integrals are analogous to spherical harmonics while more similar in structure
to the moments and products of inertia[7].
(cid:90)
(cid:104)
T l,m,n =
1
M Rl+m+n
B
xlymzndm, where l + m + n = N
(4)
A more general form of the mutual gravity potential may be found in the
2016 Hou et al. paper.
2.1. Equations of Motion
Given the form of the mutual gravity potential the equations of motion
(EOMs) for the F2BP can be generated. The inertial form of the F2BP has
12 degrees of freedom, however, using the relative dynamics it can be reduced
to 9 degrees of freedom[6].
(cid:126)X =
(cid:126)r (cid:126)θ1
(cid:126)θ2
(cid:126)r (cid:126)ω1 (cid:126)ω2
(5)
where (cid:126)r is the relative separation vector measured from the primary to sec-
ondary, the vectors (cid:126)θ1 and (cid:126)θ2 are Euler 123 angle sets defining the inertial
orientation of the primary and the orientation of the secondary relative to
the primary, the vectors (cid:126)ω1 and (cid:126)ω2 are the angular velocities corresponding
to the Euler angles such that
6
(cid:105)T
(cid:126)θi = B i(cid:126)ωi
cos θi,3
− sin θi,3
cos θi,2 sin θi,3
− sin θi,2 cos θi,3
cos θi,2 cos θi,3
sin θi,2 sin θi,3
0
0
cos θi,2
B i =
1
cos θi,2
The rotation matrix corresponding to the Euler angle sets are
C i = C ((cid:126)θi)
(6)
(7)
(8)
where the Euler angle set is the standard Euler 123 body set. Although
non-singular attitude representations are used to numerically integrate the
binary system, we use Euler angles to describe the attitude because they
have well-behaved and intuitive linearization properties for analysis later in
this paper.
The rotation matrices corresponding to the two Euler angle sets, (cid:126)θ1 and
(cid:126)θ2, are then
C A = C ((cid:126)θ1)
C B/A = C ((cid:126)θ2) = C T
AC B
(9)
(10)
Where (cid:126)θ2 is selected as the mapping from the secondary's frame into the
primary's frame for convenience of later analysis.
Following the implementation of Maciejewski, the F2BP EOMs can be
derived from Newton's second law and Euler's EOMs as [6].
(cid:16)
(cid:126)ω2 = C B/AI −1
(cid:16) C B/AI BC T
B C T
+
(cid:17)
(cid:16)
(cid:16)
(cid:16)
7
(cid:126)r = −I −1
A
I A(cid:126)ω1 ω1 + (cid:126)MA
r − 2ω1
(cid:126)r − ω1 ω1(cid:126)r − 1
m
∂U
∂(cid:126)r
(cid:126)ω1 = I −1
A
I A(cid:126)ω1 ω1 + (cid:126)MA
(cid:17)
(cid:16)
(cid:17) ·(cid:16)
(cid:17)
(cid:17)
(cid:17)(cid:17)
(cid:126)ω1 + (cid:126)ω2
B/A
C B/AI BC T
B/A
(cid:126)ω1 + (cid:126)ω2
ω1 + (cid:126)MB
C B/A
B/A + C B/AI B
−I −1
I A(cid:126)ω1 ω1 + (cid:126)MA
A
(11)
(12)
(13)
Where the (−) operator describes the skew-symmetric matrix transform
on a vector in R3
0 −f3
f3
−f2
f2
0 −f1
0
f1
f =
(14)
To fully implement the F2BP EOMs the mutual gravitational torques, (cid:126)MA
and (cid:126)MB, must be accounted for, this is accomplished by taking the partials
of the mutual gravity with respect to the relative states of the system.
(cid:126)MB = −(cid:126)α × ∂U
∂(cid:126)α
− (cid:126)β × ∂U
∂ (cid:126)β
− (cid:126)γ × ∂U
∂(cid:126)γ
(cid:126)MA = (cid:126)r × ∂U
∂ (cid:126)R
− (cid:126)MB
Where (cid:126)α, (cid:126)β, and (cid:126)γ are columns of the rotation matrix
(cid:104)
(cid:105)
C B/A =
(cid:126)α (cid:126)β (cid:126)γ
(15)
(16)
(17)
The derivation of this formulation of the gravity torques is further de-
scribed in Maciejewski[6].
We choose not to include the effects of external perturbers, such as the
Sun and planets, because their influence on the relative dynamics is small
and would be mostly averaged out of the dynamics as the asteroids repeat
their orbit.
2.2. Doubly Synchronous Equilibrium
The high dimensionality of the nonplanar form of the F2BP makes dy-
namical analysis of the general system complex and unwieldy; we thus turn to
the doubly synchronous equilibrium for use as a simplified dynamical sand-
box. For the F2BP the doubly synchronous equilibrium occurs when the
bodies are aligned with their long axes facing each other, tidally locked and
co-orbiting. Two such equilibria exist in the F2BP, an inner and outer so-
lution, however only the outer solution is stable and seen in nature[8]. For
the remainder of this paper only the outer stable equilibrium will be referred
8
IH = H •(cid:16)
(cid:104)
(cid:105)(cid:17) • H
R2I 3x3 − (cid:126)r(cid:126)r
to as the doubly synchronous equilibrium. In order to compute the doubly
synchronous equilibrium for a given system the amended potential and its
partials with respect to the degrees of freedom must be analyzed[16]. The
amended potential is defined as
E =
H 2
2IH
+ U
(18)
Where H is the angular momentum magnitude and I H is the moment of
inertia about the angular momentum axis,
(19)
the matrix ¯¯U denotes the identity matrix and • denotes the dot product.
The scalar m is the reduced mass
I A + I B + m
Where H defines the unit direction of the angular momentum .
m =
MAMB
MA + MB
cos δ
sin δ sin λ
sin δ cos λ
H =
(20)
(21)
Where δ is an offset angle from the x-axis and λ acts as a clocking angle
about the x-axis. In the equilibrium state angle δ is π/2 radians and angle
λ is 0.
Based on knowledge of the doubly synchronous equilibrium orientation
for a second degree and order gravity field, we can simplify IH to
IH = IA,z + IB,z + mr2
(22)
The subscript z denotes the z-axis moment of inertia, or the polar moment
of inertia for the body.
For a given angular momentum of the system and set of mass parame-
ters for the system, the zeroes of the partials of the amended potential with
respect to the system degrees of freedom can be used to identify an equilib-
rium.
E(cid:126)r,E (cid:126)r,E(cid:126)θ,E (cid:126)θ
,Eδ,Eλ = 0
(23)
9
The detailed formulation of the amended potential partials used to find the
extrema are
(cid:104)
∂E
∂(cid:126)r
= −H 2
I 2
H
I3x3 − H H
m
(cid:105) • (cid:126)r +
∂U
∂(cid:126)r
∂E
∂(cid:126)θ
2 H •(cid:16)
(cid:104)
∂E
∂ξ
= − H 2
2I 2
H
= − H 2
2I 2
H
∂IH
∂(cid:126)θ
∂U
∂(cid:126)θ
+
(cid:104)
I A + I B + m
R2I 3x3 − (cid:126)r(cid:126)r
(24)
(25)
(26)
(cid:105)
(cid:105)(cid:17) • ∂ H
∂ξ
Where ξ is either λ or δ
To be stable, the Hessian of the amended potential evaluated at the equi-
librium point must be positive definite.
E (cid:126)X (cid:126)Xeq > 0
(27)
Through this approach it can be shown that only the outer equilibrium point
is energetically stable[8]. We thus conclude that the outer equilibrium is the
only equilibria of interest when observing natural systems and focus on this
configuration as the doubly synchronous equilibrium of interest.
3. Planar Dynamics Analysis
To understand the system behavior at a fundamental level, we first sim-
plify it further by applying a planar and second-order assumption. These
assumptions reduce the system to four degrees of freedom, as illustrated in
Fig. 2, such that it can be described as an inertial orbit angle θ, a relative
separation magnitude R, and the phase angle of each body relative to the
separation line φ1 and φ2.
The equations of motion for this simplified system can be derived from
a Lagrangian analysis of the dynamics[8]. The Lagrangian for the planar
problem is
L =
1
2
IA,zz
φ2
1 +
1
2
IB,zz
φ2
2 +
1
2
+(IA,zz
φ1 + IB,zz
1
2
(IA,zz + IB,zz + mR2) θ2
m R2 +
φ2) θ − Upl(R, φ1, φ2)
(28)
10
Figure 2: Diagram of the planar and second-order F2BP. The system states are an inertial
orbit angle θ, a relative separation magnitude R, and the phase angle of each body relative
to the separation line φ1 and φ2.
Where we define the term Ii,jj to be the mass normalized moment of inertia
of body i about axis j.
Application of Lagrange's equation yields the following EOMs
R = θ2R − 1
m
1
1
θ =
mR2
mR2
IA,zz
mR2
IB,zz
1 +
1 +
φ1 = −(cid:16)
φ2 = −(cid:16)
(cid:26)
+
mR2
∂Upl
∂φ1
(cid:17) 1
(cid:17) 1
(cid:20)
(cid:16)¯I A
mR2
Tr
− 2
∂Upl
∂R
∂Upl
∂φ2
− 1
mR2
− 1
mR2
R θ
R
∂Upl
∂φ2
∂Upl
∂φ1
mR2
∂Upl
∂φ1
∂Upl
∂φ2
(cid:17)
(cid:16)
(cid:16)¯I B
(cid:17) − 3
(cid:16)
(cid:17) − cos 2φ2
+ Tr
+ 2
+ 2
R θ
R
R θ
R
Where Upl is the planar simplification of Eq. 2
Upl = −GMAMB
1 +
R
2R2
1
(cid:16)
+IB,xx + IB,yy − cos 2φ1
IA,yy − IA,xx
(cid:17)(cid:17)(cid:21)(cid:27)
IA,xx + IA,yy
2
IB,yy − IB,xx
(29)
(30)
(31)
(32)
(33)
Where we define ¯I i to be the mass normalized inertia tensor of body i
For this realization of the system, the conditions for the doubly syn-
chronous equilibrium can be can restated using the system states to specify
axial alignment and a circular mutual orbit[17].
11
𝑨𝑩𝑹𝝓𝟏𝝓𝟐𝜽φ1 = φ2 = φ1 = φ2 = R = 0
(34)
From inspection it is clear that Uφi under these conditions is equal to
zero, such that
(35)
Thus the necessary equilibrium rotation rate, θ∗, for an equilibrium sep-
φ1 = φ2 = θ = 0
aration, R∗ is
(cid:104)
1 +
θ∗2 =
G(MA + MB)
R∗3
(cid:104)
θ∗2 =
3
2R∗2
∂Upl
∂R
1
mR∗
(cid:16)
(cid:17)
Tr
I A
+ Tr
(cid:16)
(cid:17) − 3IA,xx − 3IB,xx
(36)
(cid:105)(cid:105)
I B
(37)
The separation still must be selected to ensure the stability of the equi-
librium as defined by the second partial of the amended potential in Eq. 27.
With the planar and second-order assumption the relative separation can
be reduced to a scalar value which allows us to identify the condition for
stability[8].
(cid:114)
R∗ >
3(MAIA,zz + MBIB,zz)
m
(38)
3.1. Linear System Manifolds
Because the doubly synchronous equilibrium is an energetically stable ar-
rangement, oscillations about the equilibrium will move along closed cycles,
referred to as center manifolds.
In their linear form each center manifold
has an associated and unique frequency, this results from the purely imag-
inary eigenvalue associated with the given center manifold[18]. Because of
the dynamical coupling of the F2BP and its relation to the mass parame-
ters, analysis of oscillations about the doubly synchronous equilibrium may
provide insight into the observability of the mass parameters. To begin this
analysis, we first must identify the eigenvalues of the dynamical system. This
12
is accomplished by computing the characteristic equation of the linearized
system. The linearized dynamics are described by
(cid:126)X = A (cid:126)X =
(39)
0
0
0
0
0
0
0
0
0
0
0
0
∂ R
∂ R
∂R 0
∂φ1
∂ θ
∂ θ
∂R 0
∂φ1
∂ φ1
∂R 0 ∂ φ1
∂R 0 ∂ φ2
∂ φ2
∂φ1
∂φ1
0
0
0
0
∂ R
∂φ2
∂ θ
∂φ2
∂ φ1
∂φ2
∂ φ2
∂φ2
1
0
0
0
0
∂ θ
∂ R
∂ φ1
∂ R
∂ φ2
∂ R
0
1
0
0
∂ R
∂ θ
∂ θ
∂ θ
∂ φ1
∂ θ
∂ φ2
∂ θ
0 0
0 0
1 0
0 1
0 0
0 0
0 0
0 0
R
θ
φ1
φ2
R
θ
φ1
φ2
The eigenvalues of the system are roots of the characteristic equation
A − λI 8x8 = 0
(40)
which can be related to a period of motion of the linearized oscillations
Pβi =
2π
Im(λi)
(41)
Here βi specifies the imaginary component of the corresponding root or
eigenvalue, λi. The described periodic behavior moves along a linear manifold
defined by a deviation vector from the equilibrium
δ (cid:126)Xλi =
Re((cid:126)ui)
Re((cid:126)ui)2
cos θi +
Im((cid:126)ui)
Im((cid:126)ui)2
sin θi
(42)
Where (cid:126)ui are the eigenvectors corresponding to a given eigenvalue.
For the planar system we have already identified one fundamental frequency
in the form of the equilibrium orbit rate constraint, θ∗. The eigen decompo-
sition provides an additional three oscillations, or fundamental frequencies.
Each is associated with one of the planar states; that is the relative separa-
tion R, and the primary and secondary phase angles, φ1 and φ2. Because
the orbit angle θ is an ignorable coordinate it has a zero eigenvalue and thus
no associated frequency from the eigen decomposition. In Table 2 the linear
periods for the 617 Patroclus system are listed.
In the linear form, the center manifolds each form an elliptic oscillation of
arbitrary amplitude about their associated state with a unique linear period
13
Table 2: Linear periods of manifolds about the planar doubly synchronous equilibrium
evaluated for the 617 Patroclus system.
Manifold Linear Period [days]
r
φ1
φ2
Orbit
3.93
13.60
12.12
4.41
(Table 2). Of interest in the linear system is the influence of the mass pa-
rameters on the linear periods of each manifold. To explore this we evaluate
the linear periods of the system as the density, volume and axes of each body
are scaled, shown in Fig. 3-5.
Figure 3: Behavior of planar doubly synchronous manifold linear periods as the density
of the system is scaled.
As the density is scaled upwards in Fig. 3 the periods each decrease in
length implying a link between the system mass and the speed of its motion.
The two phase angle periods also show very similar behavior with a roughly
constant gap between their periods, a unique behavior associated with den-
sity scaling. In the case of volumetric scaling, Fig. 4, the separation and
secondary phase angle periods show a muted response, while the behavior
of the primary phase angle period shows a high sensitivity. In this figure,
and the axis scaling figure, Fig. 5, there is an apparent switching of the pri-
mary and secondary phase angle periods near the unity scaling factor. This
14
Figure 4: Behavior of planar doubly synchronous manifold linear periods as the volume of
each body is scaled.
results from the secondary becoming more massive as one of the bodies is
scaled resulting in the gravitational dominance switching from the primary
to the secondary.
It is also of note that the scaling of the minor and in-
termediate axes is truncated near the scaling factor of one, this is done to
avoid degeneration of the system by scaling the body to a oblate spheroidal
shape.
In the axial scaling analysis, Fig. 5, as each of the three axes is
independently scaled, regardless of the body, a unique response occurs in the
periods for each axis scaling; implying unique behavior associated with each
moment of inertia. While this analysis does not provide a definite method of
determining the observability of the mass parameters, it does point towards
unique behaviors of the system as different aspects of the mass distribution
are scaled. We thus conclude that oscillations about the equilibrium likely
will impact observations of binary systems to a significant degree.
3.2. Planar Observability
To understand the influence of the mass parameters on the fundamental
frequencies, we continue with analysis of the planar and second-order F2BP.
The periods of the fundamental frequencies are considered as idealized mea-
surements made by observers. While direct observations of these frequencies
would require extensive observations in close proximity, the frequencies more
succinctly contain the same information content as direct observations of the
dynamics and help to provide a best case analysis. We approximate the
sensitivity of the fundamental frequencies to the mass parameters as
15
Figure 5: Behavior of planar doubly synchronous manifold linear periods as the length of
each ellipsoid axis is scaled.
δ(cid:126)Ω =
• δ (cid:126)T
∂(cid:126)Ω
∂ (cid:126)T
(43)
Where the vector (cid:126)Ω is the set of fundamental frequencies derived from
the eigen decomposition and the doubly synchronous orbit rate
with βi representing the linearized manifold periods, Pβi. The vector (cid:126)T is
the set of second order principal-axis inertia integrals
(44)
(45)
(cid:3)
(cid:126)Ω =(cid:2)(cid:126)Λ, θ(cid:3) , where (cid:126)Λ =(cid:2)βR, βφ1, βφ2
(cid:126)T =(cid:2)T 2,0,0
, T 0,0,2
, T 2,0,0
, T 0,2,0
A
B
A
A
, T 0,2,0
, T 0,0,2
B
B
(cid:3)
16
It is assumed that knowledge about the mass of each body is gained from
its relative separation and the reflex motion of the system about its center
of mass. We define ∂ (cid:126)Ω
as the sensitivity matrix, with which a least norm
∂ (cid:126)T
differential corrector can be used to estimate the mass parameters based on
the observed frequencies.
The three frequencies arising from the manifold analysis are not analytically
derived because of the complexity of the dynamics matrix and instead were
computed numerically. Because of this, the partials of these frequencies
could not be derived in closed form. Instead we leverage properties of the
characteristic equation
A − λiI = 0
0 = anλn
i + an−1λn−1
i + ... + a1λi + a0
(46)
(47)
Where the coefficients an represent coefficients found from the determinant
computation. We then take the partial of the characteristic equation in its
polynomial form with respect each element j of the mass parameter vector,
(cid:126)T
0 =
∂an
∂ (cid:126)T (j)
λn
i + an
∂λi
∂ (cid:126)T (j)
λn−1
i +
∂an−1
∂ (cid:126)T (j)
λn−1
i + an−1
∂λi
∂ (cid:126)T (j)
λn−2
i + ...
(48)
+
∂a1
∂ (cid:126)T (j)
λi + a1
∂λi
∂ (cid:126)T (j)
∂λi
∂ (cid:126)T (j)
by substituting the numerically
We can then solve for the the partial
generated values of λ. Using the imaginary component of this partial, we
can compute the matrix ∂ (cid:126)β
∂ (cid:126)T
for use in the sensitivity matrix.
For the planar and second-order realization of the problem the sensitivity
matrix is rank deficient at rank 4 while having 6 columns from vector (cid:126)T . As a
result the differential corrector will find a solution lying in a two dimensional
solution plane, defined by two nullspace vectors of the sensitivity matrix.
Because the mass parameters of interest are the second order principal-axis
inertia integrals, we can utilize the definition of the inertia ellipsoid
Izz ≤ Ixx + Iyy
Izz ≥ Iyy ≥ lxx
(49)
(50)
and thus constrain the valid area on the solution plane.
17
If such an estimation scheme were to be used, then other system observa-
tions and measurements would need to be gathered to find an exact solution.
For instance flybys of either asteroid measuring the spherical harmonics or
other mass tracking techniques. In the case of spherical harmonics measure-
ments, constraints can be derived from the relationship[19]
Ix − Iy = −4M r2
Iy − Iz = r2
sC22
s(C20 − 2C22)
(51)
(52)
Where rs is an arbitrary scaling length. The projection of these constraint
lines onto the solution plane provides the information necessary to reduce the
estimated solution from a two dimensional space to a single point.
We illustrate the planar estimation approach in Fig. 6 by computing
a set of initial mass parameter guesses with a Gaussian perturbation up
to 5% from the "truth" mass parameter values. These are then projected
onto the solution plane, shown as the red points, where the axes are the
two nullspace vectors. We then project the inertia ellipsoid constraints onto
the solution plane as the sets of magenta and green lines, for the primary
and secondary respectively. Finally, the spherical harmonics constraints are
projected onto the solution plane as the blue and black lines for the primary
and red and cyan lines for the secondary.Because of the similarity in shape
between Patroclus and the secondary Menoetius, the constraint lines appear
to be overlapping in the figure. For a more dissimilar binary pair the spherical
harmonics constraint lines would be more distinct. With all of these bounds
and constraints in the figure, the different steps of this estimation process
are clear; resulting in the final green truth at the crossing of the spherical
harmonics constraints.
4. Nonplanar Mass Parameter Observability
4.1. Expansion To Nonplanar Manifolds
The restrictions of the planar system enable a simple, but limited, ap-
proach to detailed analysis of the observability. However understanding the
impact of out-of-plane motion on the number of manifolds and their behav-
ior is vital for a full understanding of the mass parameter observability. We
thus expand our dynamics model to the more general model with the mutual
gravity potential truncated at order two, recalling Eq. 11-13. The increased
18
Figure 6: Projection of planar mass parameter estimates onto the solution nullspace plane.
The inertia constraints illustrate the bounds caused by the inertia ellipsoid projected onto
the plane. The C20 and C22 constraints illustrate the knowledge added by measurement
of either bodies second order spherical harmonics terms; because of the similar shape of
the two bodies simulated the spherical harmonics constraints overlap. The axes represent
unitless perturbations along each nullspace vector.
dimensionality of the nonplanar system results in seven system center mani-
folds from the eighteen states, as opposed to the three system manifolds from
eight states in the planar problem. There are also two ignorable coordinates
associated with zero-valued eigenvalues as opposed to the single ignorable
coordinate of the planar case. Because of the complexity of developing the
mutual gravity potential and torques as well as the equations of motion we
choose not to reduce the state representation to a minimal set. Because the
doubly synchronous behavior does not change from the planar to nonplanar
dynamics the θ constraint remains unchanged. The behaviors of each of these
eight fundamental frequencies are illustrated in Fig. 7.
19
Figure 7: Diagram of the nonplanar doubly synchronous equilibrium. The perturbations
associated with each system center manifold are illustrated to visualize their behavior.
We see that the three manifolds from the planar case remain for the non-
planar case, now being referred to as β3, β4, and β5. Two of the new manifolds
are associated with precession and nutation of the primary and secondary,
respectively β2 and β7. The remaining two new manifolds are associated
with a relative axial twist about the radial axis, β1, and the precession and
nutation of the orbit plane, β6. The periods of the linear manifolds for these
fundamental frequencies are shown for the 617 Patroclus system in Table 3.
As with the planar case, for observations of the nonplanar system, we
make the assumption that the reflex motion has been well characterized,
providing constraints on the relative separation and the mass ratio, defined
here as
µ =
MB
MA + MB
(53)
Thus the mass of each body can be evaluated based on an estimate of the
total mass
As a result our estimated mass parameters can be the total mass and the
MT = MA + MB
(54)
20
1𝝎1𝝎2ሶ𝜃Table 3: Linear periods of manifolds about the nonplanar doubly synchronous equilibrium
evaluated for the 617 Patroclus system.
Manifold Linear Period [days]
β1
β2
β3
β4
β5
β6
β7
Orbit
18.97
15.67
13.65
12.13
4.41
3.86
2.91
4.41
second order inertia integrals for both bodies.
(cid:126)T =(cid:2)MT , T 2,0,0
A
, T 0,2,0
A
, T 0,0,2
A
, T 2,0,0
B
, T 0,2,0
B
, T 0,0,2
B
(cid:3)
(55)
Given the increase in the number of system manifolds the system frequency
vector, (cid:126)Ω, for the nonplanar estimation becomes
(cid:126)Ω =(cid:2)β1, β2, β3, β4, β5, β6, β7, θ(cid:3)
(56)
For this estimation we now have more observables than estimated values,
thus the problem is overconstrained and can provide a full solution for the
mass parameters from a theoretical standpoint.
The process of estimation will again apply Eq. 43, however the larger
and more complex dynamics matrix for the nonplanar problem requires a
new approach to computing the partials of the fundamental frequencies with
respect to the mass parameters. Properties of the left and right eigenvectors
are leveraged to compute the frequency partials. The left eigenvectors are
defined as
λi(cid:126)vi = AT ( (cid:126)T )(cid:126)vi
while the right eigenvectors are defined as
λi(cid:126)ui = A( (cid:126)T )(cid:126)ui
(57)
(58)
where the i indicates the specific eigenvalue-vector pair. To begin the deriva-
tion, the partial of the right eigenvalue equation is taken with respect to the
21
jth mass parameter
∂λi
∂ (cid:126)T (j)
(cid:126)ui + λi
∂(cid:126)ui
∂ (cid:126)T (j)
=
∂A
∂ (cid:126)T (j)
(cid:126)ui + A
∂(cid:126)ui
∂ (cid:126)T (j)
(59)
Left multiplying this partial by the transpose of the left eigenvectors, (cid:126)vT
equation becomes
i , the
∂λi
∂ (cid:126)T (j)
i (cid:126)ui + λi(cid:126)vT
(cid:126)vT
i
∂(cid:126)ui
∂ (cid:126)T (j)
= (cid:126)vT
i
∂A
∂ (cid:126)T (j)
(cid:126)ui + (cid:126)vT
i A
∂(cid:126)ui
∂ (cid:126)T (j)
(60)
in which the ∂(cid:126)ui
∂ (cid:126)T (j)
Rearranging to solve for the partial of the frequency, βi only
terms cancel based on the definition of the left eigenvector.
(cid:16) 1
(cid:17)
∂βi
∂ (cid:126)T (j)
= Im
(cid:126)vT
i
∂A
∂ (cid:126)T (j)
(cid:126)ui
(cid:126)vT
i (cid:126)ui
∂(cid:126)Ω
∂ (cid:126)P
=
diag(cid:0)(cid:126)Ω(cid:1)2
2π
22
Thus the sensitivity matrix for the nonplanar problem can be computed
element by element, iterating over this equation.
Given the complete solution generated by the estimation process, we can
now analyze the uncertainties of the estimated mass parameters. The covari-
ance and correlation of the mass parameters generated from this approach
help to quantify the influence of the mass parameters on dynamical observa-
tions. To begin this analysis the covariance matrix for the mass parameters,
PT T , is formulated by relating the pseudo-inverse of the sensitivity matrix
and the observational covariance of the fundamental frequencies, P ΩΩ
(cid:17)−1 ∂(cid:126)Ω
(cid:16) ∂(cid:126)Ω
∂ (cid:126)T
T
∂(cid:126)Ω
∂ (cid:126)T
T
• P ΩΩ • ∂(cid:126)Ω
∂ (cid:126)T
∂ (cid:126)T
(cid:17)−1
(cid:16) ∂(cid:126)Ω
∂ (cid:126)T
T
∂(cid:126)Ω
∂ (cid:126)T
P T T = δ (cid:126)T δ (cid:126)T T =
(62)
Because the frequencies are not an intuitive measurement the covariance of
the fundamental frequencies is converted to the covariance of fundamental
periods
P ΩΩ =
T
∂(cid:126)Ω
∂ (cid:126)P
• P P P • ∂(cid:126)Ω
∂ (cid:126)P
(63)
This conversion is simply the derivative of the frequency and period relation-
ship.
(61)
(64)
Table 4: Allowable observational variance of observations for 10% certainty of mass pa-
rameter estimates of the 617 Patroclus system.
Mass Parameter Observational Accuracy Requirement [sec]
A
MT
T 200
T 020
T 002
T 200
T 020
T 002
B
B
A
A
B
10584.0
25.3
25.3
25.3
15.0
15.0
15.0
To define the covariance of the fundamental periods it is assumed that all
periods would be independently measured using the same observation tech-
nique.
P P P = σ2
P • I
(65)
where there is a single period observational variance, σP , that is applied to the
observations of each period. This observational variance can be considered
to be a temporal resolution of the period measurements or the precision of
each measurement.
4.2. An Idealized Estimation Method
This formulation of the covariance allows us to treat the sensitivity anal-
ysis approach as an idealized estimation approach. Here the observational
variance, σP , would act as an observational accuracy requirement to constrain
the mass parameter estimates to a specific accuracy level. In Table 4 this
is leveraged to predict the observational variance necessary in order to gain
10% knowledge of each of the seven mass parameters of the 617 Patroclus
system.
To provide further insight, we compute the mass parameter variance for
617 Patroclus, given a 1 second observation variance
¯σT =(cid:2)9.45 × 10−6 3.95 × 10−3 3.94 × 10−3 3.95 × 10−3
6.64 × 10−3 6.64 × 10−3 6.64 × 10−3(cid:3)T
(66)
23
and the correlation matrix
(67)
1.
ρT T =
1.
0.99
0.99
0.99
1.
0.99
−0.27 −0.27 −0.27 −0.64 −0.64 −0.64
0.99 −0.55 −0.55 −0.55
0.99 −0.55 −0.55 −0.55
−0.55 −0.55 −0.55
1.
0.99
0.99
1.
−0.27
−0.27
−0.27
−0.64 −0.55 −0.55 −0.55
−0.64 −0.55 −0.55 −0.55
−0.64 −0.55 −0.55 −0.55
1.
0.99
0.99
0.99
1.
0.99
The values for the variance and correlation in Eq. 66 and 67 are generated
from the covariance matrix, which is not included for the sake of brevity
and clarity. The bar used with variance values are normalized by the mass
parameters corresponding to each element, such that they represent fractional
covariances and variances
¯P T T (i, j) =
¯σT (i) =
ρT T (i, j) =
PT T (i, j)
(cid:126)T (i) (cid:126)T (j)
σT (i)
(cid:126)T (i)
¯P T T (i, j)
¯σT (i)¯σT (j)
(68)
(69)
(70)
4.3. Mass Parameter Scaling Effects on Observability
While these results show the mass parameters to be well estimated based
on their variances, the observational requirement of 1 second variance on each
frequency is highly restrictive. Likewise, the observational requirements to
achieve 10% knowledge of the inertia integrals in table 4 further illustrates
the accuracy of observations necessary for second order parameter measure-
ments. In combination with the high correlation on the second order param-
eters, it becomes clear that without further constraints, via in-situ gravity
measurement or other methods, such an estimation approach is not feasible
even with these idealized observations.
To better understand the observability of binary systems as a whole we
now investigate the effects of scaling the mass parameters on the correlation
and covariance matrices. Specifically we scale the mass ratio and second order
principal inertia integrals of Menoetius, the 617 Patroclus secondary. This is
done as a means to explore the effects of mass ratio and differing asteroid mass
24
distributions on the mass parameter observability. We do not simultaneously
scale the second order asteroid inertia integrals as this would merely scale the
correlation and covariance as opposed to changing their structure. Likewise
only the secondary is scaled as the system is nearly symmetric such that
scaling either body will show the same observability behavior. The scaling
of the inertia integrals is limited such that the asteroid remains a triaxial
ellipsoid as opposed to an oblate spheroid; this is to avoid the degeneration
of the manifolds of the system caused by symmetric semi-axes.
For simplicity the second order principal axis inertia integrals of the sec-
ondary are refered to here as Tx, Ty, and Tz. The study performed scales
the mass ratio by .1 and .55, Tx by .88 and .94, Ty by .7 and .85, and Tz
by .01 and .51. The scaling floor for Tx and Ty is selected to avoid the the
inertia integral dropping below the next smallest inertia integral which would
cause an spherical degeneracy, thus Tz has no floor. Fig. 8 and 9 illustrate
the covariance and correlation with the covariance elements colored by their
log value and the correlation elements colored by their linear value from -1
to 1. The key point of interest in this case is whether the scaling of these
mass parameters can lower the correlation between the each body's second
order inertia integrals. What both results show is that outside of the case
of an extremely flattened body, the second order inertia integrals remain
highly correlated. Even in the case of an extremely flattened body only the
Tz correlation changes significantly while the Tx and Ty relationship remains
very coupled. This implies that the dynamical effects of these unknown mass
parameters can be significant for most shapes and configurations.
5. Near-Spherical Doubly Synchronous Systems
For a number of observed doubly synchronous binaries current shape
knowledge is limited to a mean radius value or spherical shape estimate.
However, for a binary system to remain in a stable doubly synchronous or-
bit, dynamical analysis has shown that the mass distribution of the bodies
must be elongated such that the bodies mutual gravity torques exist to en-
force the tidal locking. One result of this is that for a sphere-sphere doubly
synchronous system four of the manifolds become zero eigenvalues due to the
lack of attitude interaction between the bodies, described in Table 5 for the
Pluto-Charon system. In addition, for a nearly spherical body the effects of
any elongation will be so low that the periods of these four manifolds will be
functionally immeasurable.
25
Figure 8: Effects of secondary mass parameter scaling on mass parameter covariance
matrix with magnitude of matrix elements shown in log color.
To better analyze these near spherical systems we perform a linearization
about the spherical shape of the bodies using an elongation factor which
perturbs the bodies as triaxial ellipsoids defined such that
a = R(1 + ),
abc = R3
b = R,
Ix =
2M R2
5
(1 − ),
Iy =
2M R2
5
26
c = R(1 − )
2M R2
5
,
Ix =
(71)
(72)
(73)
(1 + )
246*=.1246-10-9-8-7-6-5Mass Ratio Scaling Covariance246*=.55246-9-8-7-6-5-4246*=s246-10-9-8-7-6-5246T*x=.88Tx246-9-8-7-6-5Secondary Tx Scaling Covariance246T*x=.94Tx246-10-8-6-4246T*x=Tx246-10-9-8-7-6-5246T*y=.7Ty246-10-9-8-7-6-5Secondary Ty Scaling Covariance246T*y=.85Ty246-10-9-8-7-6-5246T*y=Ty246-10-9-8-7-6-5246T*z=.01Tz246-10-9-8-7-6-5Secondary Tz Scaling Covariance246T*z=.51Tz246-10-9-8-7-6-5246T*z=Tz246-10-9-8-7-6-5Figure 9: Effects of secondary mass parameter scaling on mass parameter correlation
matrix with magnitude of matrix elements shown in linear color.
where a, b, and c are the semi-axes.
Using these definitions of the near-spherical mass distributions we can
linearize the dynamics as
δ(cid:126)Ω =
(cid:104) ∂(cid:126)Ω
(cid:105)
δ(cid:126)
(74)
∂(cid:126)
sphere
27
246*=.1246-0.500.51Mass Ratio Scaling Correlation2468*=.552468-0.500.51246*=246-0.500.51246T*x=.88Tx246-0.500.51Secondary Tx Scaling Correlation246T*x=.94Tx246-0.500.51246T*x=Tx246-0.500.51246T*y=.7Ty246-0.500.51Secondary Ty Scaling Correlation246T*y=.85Ty246-0.500.51246T*y=Ty246-0.500.51246T*z=.01Tz246-0.500.51Secondary Tz Scaling Correlation246T*z=.51Tz246-0.500.51246T*z=Tz246-0.500.51(cid:104) ∂(cid:126)Ω
∂ (cid:126)T
(cid:105)
∂ (cid:126)T
∂(cid:126)
δ(cid:126)Ω =
δ(cid:126)
sphere
∂ (cid:126)T
∂(cid:126)
=
−2R2
P
5
0
0
−2R2
C
5
0
0
0
0
2R2
P
5
0
0
2R2
C
5
(75)
(76)
where (cid:126)Ω represents the remaining system frequenciess, (cid:126)T again represents
the vector of only the second-order principal-axis inertia integrals, and (cid:126) rep-
resents the elongation factors applied to the primary and secondary body
independently. It is of note that this linearization assumes that the equilib-
rium separation remains constant as the bodies are elongated. This means
that the equilibrium orbit rate must scale as the bodies are elongated while
maintaining the observed separation. This effect on the orbit rate is analo-
gous to the effect on the other three periods of the elongation, however they
are not as simply expressed as the orbit rate, whose change can be directly
computed. By analyzing the linearized effects of the deformation on the four
remaining periods and observing the behavior of these measurable manifold
periods, we provide a different approach to understand the effects of the mass
distributions of near-spherical systems.
5.1. Application to Pluto-Charon System
To illustrate the use of this approach we apply it to the Pluto-Charon
system, a doubly synchronous binary for which only mean radius information
has been reliably measured. The density and shape results of Nimmo et al.'s
analysis of New Horizons images report the density of Pluto and Charon
m3 and the mean radii to be 1188.3±1.6 km and
to be 1854 kg
606.0±1.0 km respectively[20]. Applying our analysis to these parameters
the system periods can be computed for the spherical system, Table 5.
m3 and 1701 kg
Beginning from the spherical system periods we apply the linearization
in three ways to understand what information can be gained from this anal-
ysis. The first approach is to perturb only the shape of Pluto, next only
the shape of Charon is perturbed, and finally the shape of both bodies are
28
Table 5: Linear periods of manifolds about the nonplanar doubly synchronous equilibrium
evaluated for the spherical Pluto-Charon system.
Manifold Linear Period [days]
β1
β2
β3
β4
β5
β6
β7
Orbit
DNE
DNE
DNE
DNE
6.39
6.39
6.39
6.39
identically perturbed. This is to say that the vector (cid:126) can be expressed as
[1, 0] · P luto, [0, 1] · Charon, and [1, 1] · system respectively. From these three
approaches, illustrated in Fig. 10, we see that the relative behavior of the pe-
riods differs uniquely for deformation of each body. This implies that through
measurement of the relative lengths of the periods, information on the mass
distribution could be gathered to help constrain each body's mass distribu-
tion. As is clear from these results however, the deformation of the bodies
would need to be sufficiently large to be detected by realistic measurement
methods.
For a comparison point we also compute the complete set of nonplanar
periods associated with an ellipsoidal mass distribution for both Pluto and
Charon generated with a value of =.0008. This value of is selected because
it lies in the middle of the certainty bounds on the mean radius values re-
ported by Nimmo et al; representing roughly a 1 km deformation in Pluto's
semi-axes and a .5 km change in Charon's semi-axes. This is reported in Ta-
ble 6 and further confirms the difficulty of applying our approach to a system
so near to the spherical case. The four short periods would not be feasible
to distinguish or measure with sufficient accuracy. The four long periods on
the other hand would require logistically impossible measurement efforts to
be accurately and precisely observed, due to their length.
6. Conclusions and Future Work
In this paper we have shown that the influence of mass parameters on the
observable dynamics of binary asteroids, up to the second order, in the dou-
29
Figure 10: Linearized change in Pluto-Charon manifold periods as shape is modified by
parameter. From left to right the figures show and elongation of only Pluto, elongation of
only Charon, and equal elongation of both Pluto and Charon
bly synchronous equilibrium will affect dynamical observations significantly.
The use of the doubly synchronous assumption allowed for a relatively sim-
ple differential corrector method to elucidate fundamental frequencies of the
system as a target binary oscillates near the equilibrium. In order to accom-
plish this we analyzed the manifolds of the planar problem and investigated
the influence of mass parameters on the linear behavior of these manifolds
to understand how they may affect observations of these systems. For the
planar case we found that the mass parameters were not fully observable
based purely on observations of the dynamics and would require other in-
situ or remote observations to constrain a system's mass parameters. For
the nonplanar F2BP we were able to show that the mass parameters were
fully observable using only observations of the system dynamics, although
the observational requirements are demanding. For the nonplanar differen-
30
Table 6: Linear periods of manifolds about the nonplanar doubly synchronous equilibrium
evaluated for the =.0008 Pluto-Charon system.
Manifold Linear Period [days]
β1
β2
β3
β4
β5
β6
β7
Orbit
4659.25
1442.06
393.35
96.37
6.39
6.39
6.33
6.39
tial corrector estimation we were able to investigate the achievable covariance
of the estimated mass parameters based on the accuracy of observations of
the system dynamics. This provides an understanding of the information
quality requirements for the proposed mass parameter estimation approach
to be effective. From this analysis we can conclude that more robust mea-
surements, likely from an in-situ spacecraft, would be necessary for mass
parameter estimation. Finally, a limited approach to the application of this
analysis to near-spherical systems was presented and applied to the Pluto-
Charon system.
Appendix A. The Hou Mutual Gravity Potential
The general order Hou et al. reformulation of the mutual gravity potential
begins from the double integral description of the mutual gravity potential.
N(cid:88)
U = −G
1
Rn+1
n=0
Un
(A.1)
A binomial expansion and Legendre polynomial expansion are then per-
formed to approximate the mass distribution to order N with the inertia
31
integrals in a recursive summation.
(cid:88)
0(cid:88)
Un =
tn
k
×bn−k
ak
(i1,i2,i3)(i4,i5,i6)
(A.2)
k(2)=n
(i1,i2,i3)(i41,i5,i6)(j1,j2,j3)(j4,j5,j6)
(j1,j2,j3)(j4,j5,j6)ei1+i4
×MBT
x
ei2+i5
y
(cid:48)(i4+j4),(i5+j5),(i6+j6)
B
A
z MAT (i1+j1),(i2+j2),(i3+j3)
ei3+i6
Where k(2) implies stepping by 2 as opposed to 1
Here the recursive coefficient is represented by tn
k .
k+2 = −(n − k)(n + k + 1)
(k + 2)(k + 1)
tn
(A.3)
While the binomial expansion coefficients are ak
(i1,i2,i3)(i4,i5,i6) and bn−k
(j1,j2,j3)(j4,j5,j6).
(A.4)
(A.5)
(i1,i2,i3)(i4,i5,i6) = ak−1
ak
(i1,i2,i3−1)(i4,i5,i6) − ak−1
+ak−1
−ak−1
(i1−1,i2,i3)(i4,i5,i6) + ak−1
(i1,i2,i3)(i4−1,i5,i6) − ak−1
(i1,i2,i3)(i4,i5,i6−1)
(i1,i2−1,i3)(i4,i5,i6)
(i1,i2,i3)(i4,i5−1,i6)
(j1,j2,j3)(j4,j5,j6) = bk−2
bk
(j1−2,j2,j3)(j4,j5,j6) + bk−2
+bk−2
(j1,j2,j3−2)(j4,j5,j6) + bk−2
−2bk−2
(j1−1,j2,j3)(j4−1,j5,j6) − 2bk−2
(j1,j2,j3)(j4−2,j5,j6) + bk−2
(j1,j2−1,j3)(j4,j5−1,j6) − 2bk−2
(j1,j2−2,j3)(j4,j5,j6)
(j1,j2,j3)(j4,j5−2,j6) + bk−2
(j1,j2,j3)(j4,j5,j6−2)
(j1,j2,j3−1)(j4,j5,j6−1)
For these coefficients the superscripts and subscripts serve as indices. The i
and j expansion indices are constrained by and summation over these indices
sums over the possible combinations of these indices based on the values of
k and n and the constraint equations.
k = i1 + i2 + i3 + i4 + i5 + i6
n − k = j1 + j2 + j3 + j4 + j5 + j6
(A.6)
(A.7)
The variables R and ex, ey, and ez represent the magnitude and unit direction
of the relative separation. The variables TA and T (cid:48)
B are the mass-normalized
inertia integral sets for the primary and secondary bodies, where the prime
denotes that the inertia integrals of the secondary are rotated into the frame
of the primary.
32
Appendix A.1. Inertia Integrals
Central to this reformulation of the mutual potential is the use of inertia
integrals to describe the mass distribution. This aspect of the reformulation
is accomplished by the application of a Legendre polynomial expansion to
describe the mass distributions, where the Legendre coefficients are referred
to as inertia integrals.
T l,m,n =
1
M Rl+m+n
B
xlymzndm, where l + m + n = N
(A.8)
In this way the inertia integrals can be considered analogous in use to spher-
ical harmonics[7]. The mathematical form of the inertia integrals is similar
to that of the moments and products of inertia for a rigid body wherein each
term represents the mass distribution about some axis, however the inertia
integrals are expanded to order N whereas moments of inertia are linear
combinations of second order inertia integrals. Here we provide the mass
and length normalized form of an inertia integral along with the 0th and 2nd
order coefficients in terms of the normalized moments and product of inertia.
(cid:90)
(cid:90)
T l,m,n =
1
M Rl+m+n
B
xlymzndm, where l + m + n = N
1 = T 0,0,0
Ixx = T 0,2,0 + T 0,0,2
Iyy = T 2,0,0 + T 0,0,2
Izz = T 2,0,0 + T 0,2,0
Ixy = −T 1,1,0
Ixz = −T 1,0,1
Iyz = −T 0,1,1
(A.9)
(A.10)
(A.11)
(A.12)
(A.13)
(A.14)
(A.15)
(A.16)
It is of note that the 0th order inertia integral is equal to the mass for the
non-normalized form, thus it is equal to one in the normalized form.
Appendix B. Planar Dynamics Matrix
(cid:126)X =(cid:2)r θ φ1 φ2
(cid:3)T
r
θ
φ1
φ2
(B.1)
33
0 0
0 0
1 0
0 1
0 0
0 0
0 0
0 0
0
0
0
0
∂ r
∂r
∂ θ
∂r
∂ φ1
∂r
∂ φ2
∂r
0
0
0
0
0
0
0
0
∂ r
0
∂φ1
∂ θ
0
∂φ1
0 ∂ φ1
∂φ1
0 ∂ φ2
∂φ1
0
0
0
0
∂ r
∂φ2
∂ θ
∂φ2
∂ φ1
∂φ2
∂ φ2
∂φ2
1
0
0
0
0
∂ θ
∂ r
∂ φ1
∂ r
∂ φ2
∂ r
0
1
0
0
∂ r
∂ θ
∂ θ
∂ θ
∂ φ1
∂ θ
∂ φ2
∂ θ
(cid:126)X = A (cid:126)X =
∂r
∂r
= θ2 − Vrr
m
∂r
∂φ1
= −Vrφ1
m
∂r
∂φ2
= −Vrφ2
m
∂r
∂ θ
= 2 θr
∂ θ
∂r
= −2
Vφ1
mr3 +
Vφ1
mr2 − 2
Vφ2
mr3 +
Vφ2
mr2 + 2
r θ
r2
∂ θ
∂φ1
∂ θ
∂φ2
=
Vφ1φ1
mr2
=
Vφ2φ2
mr2
∂ θ
∂ r
∂ θ
∂ θ
= −2
= −2
θ
r
r
r
34
r
θ
φ1
φ2
r
θ
φ1
φ2
(B.2)
(B.3)
(B.4)
(B.5)
(B.6)
(B.7)
(B.8)
(B.9)
(B.10)
(B.11)
(cid:0)Vφ1 + Vφ2
(cid:1)
mr3
−
(cid:0)Vrφ1 + Vrφ2
(cid:1)
mr2
∂ φ1
∂r
= 2
− Vrφ1
MAIA,zz
− 2
r θ
r2
∂ φ1
∂φ1
= −Vφ1φ1
mr2 − Vφ1φ1
MAIA,zz
∂ φ1
∂φ2
= −Vφ2φ2
mr2
∂ φ1
∂ r
= 2
θ
r
(cid:0)Vφ1 + Vφ2
(cid:1)
mr3
−
∂ φ2
∂r
= 2
∂ φ1
∂ θ
= 2
r
r
(cid:0)Vrφ1 + Vrφ2
mr2
(cid:1)
− Vrφ2
MBIB,zz
− 2
r θ
r2
∂ φ2
∂φ1
= −Vφ1φ1
mr2
∂ φ2
∂φ2
= −Vφ2φ2
mr2 − Vφ2φ2
MBIB,zz
∂ φ2
∂ r
∂ φ2
∂ θ
= 2
= 2
θ
r
r
r
35
(B.12)
(B.13)
(B.14)
(B.15)
(B.16)
(B.17)
(B.18)
(B.19)
(B.20)
(B.21)
GMAMB
Vr =
r2
−3
2
Vrr = 2
GMAMB
GMAMB
− 6
r3
r5
−3
2
For second order inertia tensor formulation of mutual potential, partials
of the potential are as follows:
(B.23)
(B.24)
(B.25)
(B.22)
3
2r2
Vφ2 = 3
Vφ1 = 3
(cid:1)
(cid:1)
GMAMB
2r3
GMAMB
2r3
(cid:0)IA,xx + IA,yy + IA,zz + IB,xx + IB,yy + IB,zz
(cid:0)1 +
(cid:1)(cid:0)IA,yy − IA,xx
(cid:1) + IB,xx + IB,yy
(cid:0)IA,xx + IA,yy − cos(cid:0)2φ1
(cid:1)(cid:1)(cid:1)(cid:1)
(cid:1)(cid:0)IB,yy − IB,xx
−cos(cid:0)2φ2
(cid:1)(cid:0)IA,yy − IA,xx
sin(cid:0)2φ1
sin(cid:0)2φ2
(cid:1)(cid:0)IB,yy − IB,xx
(cid:0)IA,xx + IA,yy + IA,zz + IB,xx + IB,yy + IB,zz
(cid:1)(cid:0)IA,yy − IA,xx
(cid:1) + IB,xx + IB,yy
(cid:1)(cid:1)(cid:1)
(cid:1)(cid:0)IB,yy − IB,xx
sin(cid:0)2φ1
(cid:1)(cid:0)IA,yy − IA,xx
(cid:1)
(cid:1)
(cid:1)(cid:0)IB,yy − IB,xx
sin(cid:0)2φ2
(cid:1)
(cid:1)(cid:0)IA,yy − IA,xx
cos(cid:0)2φ1
(cid:1)
(cid:1)(cid:0)IB,yy − IB,xx
cos(cid:0)2φ2
(cid:0)IA,xx + IA,yy − cos(cid:0)2φ1
−cos(cid:0)2φ2
Vrφ1 = −9
Vrφ2 = −9
Vφ1φ1 = 3
GMAMB
2r4
GMAMB
2r4
GMAMB
GMAMB
Vφ2φ2 = 3
r3
r3
(B.26)
(B.27)
(B.28)
(B.29)
36
Appendix C. Nonplanar Dynamics Matrix
Within this section we add the notation
to denote a skew-symmetric
matrix operator in addition to the previous tilde notation.
(cid:16)−(cid:17)s
03
03
03
(cid:126)R
∂
∂(cid:126)r
∂ (cid:126)ω1
∂(cid:126)r
∂ (cid:126)ω2
∂(cid:126)r
03
(cid:126)θ1
∂
∂(cid:126)θ1
03
03
03
03
03
03
(cid:126)θ2
∂
∂(cid:126)θ2
∂ (cid:126)r
∂(cid:126)θ2
∂ (cid:126)ω1
∂(cid:126)θ2
∂ (cid:126)ω2
∂(cid:126)θ2
03
03
03 B1
03
03
03
∂ (cid:126)r
∂ (cid:126)r
03
03
03 B2
∂ (cid:126)r
∂(cid:126)ω1
∂ (cid:126)ω1
∂(cid:126)ω1
∂ (cid:126)ω2
∂(cid:126)ω1
03
03
∂ (cid:126)ω2
∂(cid:126)ω2
(cid:17)s ∂(cid:126)γ
∂(cid:126)θ2
(cid:126)β
= −α
∂2U
∂(cid:126)α∂(cid:126)r
+
∂(cid:126)α
∂(cid:126)θ2
(cid:16)∂U
(cid:17)s ∂(cid:126)α
(cid:16)(cid:126)α
(cid:16)(cid:126)α
(cid:16)(cid:126)α
= −(cid:16) ∂U
∂2U
∂(cid:126)α2 +
∂2U
∂ (cid:126)β∂(cid:126)α
∂2U
∂(cid:126)γ∂(cid:126)α
+
+
−(cid:126)α
− (cid:126)β
−(cid:126)γ
∂(cid:126)r
(cid:17)s
− β
(cid:16)∂U
∂2U
∂ (cid:126)β∂(cid:126)r
(cid:17)s ∂ (cid:126)β
∂(cid:126)θ2
+ (cid:126)γ
+ (cid:126)γ
∂ (cid:126)β
∂2U
∂(cid:126)α∂ (cid:126)β
∂2U
(cid:126)β
∂ (cid:126)β2
∂2U
∂(cid:126)γ∂ (cid:126)β
(cid:126)β
+ (cid:126)γ
− γ
∂2U
∂(cid:126)γ∂(cid:126)r
∂(cid:126)γ
(cid:16)∂U
(cid:17)
(cid:17)
(cid:17)
+
∂2U
∂(cid:126)α∂(cid:126)γ
∂2U
∂ (cid:126)β∂(cid:126)γ
∂2U
∂(cid:126)γ2
+ r
∂2U
∂(cid:126)r2 − ∂ (cid:126)MB
∂(cid:126)r
(cid:17) − ∂ (cid:126)MB
∂(cid:126)θ2
α
∂2U
∂(cid:126)α∂(cid:126)r
+ β
∂2U
∂ (cid:126)β∂(cid:126)r
+ γ
∂2U
∂(cid:126)γ∂(cid:126)r
(cid:126)θ1
∂
∂(cid:126)θ1
=
∂B1
∂(cid:126)θ1
(cid:126)ω1
37
A =
∂ (cid:126)MB
∂(cid:126)r
∂ (cid:126)MB
∂(cid:126)θ2
=
∂ (cid:126)MA
∂(cid:126)r
(cid:16)
= r
∂ (cid:126)MA
∂(cid:126)θ2
(C.1)
(C.2)
(C.3)
(C.4)
(C.5)
(C.6)
−(cid:16)
I −1
A
(cid:16)
∂(cid:126)r
∂(cid:126)r
= rI −1
A
∂ (cid:126)MA
∂(cid:126)r
∂(cid:126)r
∂(cid:126)θ2
= rI −1
A
∂ (cid:126)MA
∂(cid:126)θ2
−(cid:16)
(cid:126)θ2
∂
∂(cid:126)θ2
=
∂B2
∂(cid:126)θ2
(cid:126)ω2
I A(cid:126)ω1 ω1 + MA
(cid:17)(cid:17)s − ω1 ω1 − 1
m
α
∂2U
∂(cid:126)α∂(cid:126)r
+ β
∂2U
∂ (cid:126)β∂(cid:126)r
+ γ
∂2U
∂(cid:126)γ∂(cid:126)r
∂2U
∂(cid:126)r2
(cid:17)
(C.7)
(C.8)
(C.9)
(C.10)
= −2ω1
∂(cid:126)r
∂ (cid:126)r
(cid:16)(cid:16)
I A(cid:126)ω1
∂(cid:126)r
∂(cid:126)ω1
= rI −1
A
(cid:17)s − ω1I A
(cid:17)
(cid:17)s
(cid:16)
ω1(cid:126)r
+ 2 r +
+ ω1r
(C.11)
∂ (cid:126)ω1
∂(cid:126)r
∂ (cid:126)ω1
∂(cid:126)θ2
= I −1
A
∂ (cid:126)MA
∂(cid:126)r
= I −1
A
∂ (cid:126)MA
∂(cid:126)θ2
(cid:16)
∂ (cid:126)ω1
∂(cid:126)ω1
= I −1
A
I A(cid:126)ω1 ω1I A
(cid:17)s
(cid:16)
(cid:126)ω1 + (cid:126)ω2
(cid:17)
ω1
∂ (cid:126)ω2
∂(cid:126)θ2
= I B
∂ (cid:126)ω2
∂(cid:126)r
∂I −1
∂(cid:126)θ2
B
= I −1
B
+ I −1
B
−(cid:16)
(cid:16)
∂ (cid:126)MA
∂(cid:126)θ2
− I B
(cid:126)ω1 + (cid:126)ω2
(cid:126)ω1 + (cid:126)ω2
38
+ (cid:126)MB
∂I −1
∂(cid:126)θ2
B
(C.12)
(C.13)
(C.14)
(C.15)
+
(C.16)
(cid:17) ∂I B
∂(cid:126)θ2
∂ (cid:126)MB
∂(cid:126)r
(cid:16) − ω1
(cid:16)
(cid:17) ∂ I B
(cid:17)∂I −1
∂(cid:126)θ2
B
∂(cid:126)θ2
(cid:126)ω1 + (cid:126)ω2
(cid:17)
− I −1
A
∂ (cid:126)MA
∂(cid:126)θ2
(cid:16)
I B
∂ (cid:126)ω2
∂(cid:126)ω1
= I −1
B
(cid:16)
(cid:126)ω1 + (cid:126)ω2
(cid:17)s(cid:17)
I A(cid:126)ω1
(cid:17)s − ω1I B − I B
(cid:16) − ω1I B − I B +
(cid:17)
+ I −1
A
(cid:16)
(cid:16)
(cid:126)ω1 + (cid:126)ω2
ω1I A −(cid:16)
(cid:17)
(cid:17)∂ I B
∂(cid:126)ω2
∂ (cid:126)ω2
∂(cid:126)ω2
= I −1
B
Appendix C.1. Total Mass Partials of Dynamics Matrix
Partials of A with respect to MT
−(cid:16)∂I −1
A
∂MT
= r
∂2(cid:126)r
∂(cid:126)r∂MT
I A(cid:126)ω1 ω1 + I −1
∂I −1
A
∂MT
∂I A
∂MT
MT (µ − µ2)
+
1
A
∂ (cid:126)MA
∂(cid:126)r
(cid:126)ω1 ω1
A
+ I −1
∂I −1
A
∂MT
∂2 (cid:126)MA
∂(cid:126)r∂MT
(cid:126)MA + I −1
∂3U
A
(cid:17)s
∂ (cid:126)MA
∂MT
∂2U
∂(cid:126)r2 − 1
m
∂(cid:126)r2∂MT
(C.17)
(C.18)
(C.19)
(C.20)
(cid:16)∂I −1
A
∂MT
∂2(cid:126)r
∂(cid:126)θ2∂MT
= r
1
+
MT (µ − µ2)
(cid:17)
∂ (cid:126)MA
∂(cid:126)θ2
+ I −1
A
∂2 (cid:126)MA
∂(cid:126)θ2∂MT
∂3U
∂2U
∂(cid:126)r∂(cid:126)θ2
+ I −1
A
(cid:17)s
− 1
m
(cid:16) ∂I −1
(cid:17)
A
∂MT
−I −1
A ω1
∂I A
∂MT
∂(cid:126)r∂(cid:126)θ2∂MT
(cid:17)s − ∂I −1
A
∂MT
(cid:126)ω1
(cid:16)
(cid:16)∂I −1
A
∂MT
= r
I A(cid:126)ω1
∂2(cid:126)r
∂(cid:126)ω1∂MT
∂2 (cid:126)ω1
∂(cid:126)r∂MT
∂2 (cid:126)ω1
∂(cid:126)θ2∂MT
=
=
∂I −1
A
∂MT
∂ (cid:126)MA
∂(cid:126)r
+ I −1
A
∂2 (cid:126)MA
∂(cid:126)r∂MT
∂I −1
A
∂MT
∂ (cid:126)MA
∂(cid:126)θ2
+ I −1
A
∂2 (cid:126)MA
∂(cid:126)θ2∂MT
39
ω1I A
(C.21)
(C.22)
(C.23)
∂2 (cid:126)ω1
∂(cid:126)ω1∂MT
=
∂I −1
A
∂MT
−∂I −1
A
∂MT
(cid:16)
I A(cid:126)ω1
(cid:17)s
+ I −1
A
(cid:16) ∂I A
∂MT
(cid:17)s
(cid:126)ω1
ω1I A − I −1
A ω1
∂I A
∂MT
∂2 (cid:126)ω2
∂(cid:126)r∂MT
=
∂I −1
B
∂MT
∂ (cid:126)MB
∂(cid:126)r
+ I −1
B
∂2 (cid:126)MB
∂(cid:126)r∂MT
− ∂I −1
A
∂MT
∂ (cid:126)MA
∂(cid:126)r
− I −1
A
∂2 (cid:126)MA
∂(cid:126)r∂MT
∂2 (cid:126)ω2
∂(cid:126)θ2∂MT
=
∂I B
∂MT
((cid:126)ω1 + (cid:126)ω2)ω1
∂I −1
∂(cid:126)θ2
B
∂2I B
∂(cid:126)θ2∂MT
+
∂2 (cid:126)MB
∂(cid:126)θ2∂MT
ω1((cid:126)ω1 + (cid:126)ω2)
− ∂I B
∂MT
−∂I −1
B
∂MT
∂ (cid:126)MB
∂(cid:126)θ2
− I B((cid:126)ω1 + (cid:126)ω2)
B
− I −1
∂2I −1
∂(cid:126)θ2∂MT
B
+ I B((cid:126)ω1 + (cid:126)ω2)ω1
B
∂I −1
∂ (cid:126)MB
∂(cid:126)θ2
∂MT
− ∂ I B
∂MT
∂ (cid:126)MA
∂(cid:126)θ2
+ (cid:126)MB
B
B
∂2I −1
∂(cid:126)θ2∂MT
∂2I −1
∂(cid:126)θ2∂MT
∂I −1
∂(cid:126)θ2
∂2 (cid:126)MA
∂(cid:126)θ2∂MT
∂2 I B
∂(cid:126)θ2∂MT
B
A
((cid:126)ω1 + (cid:126)ω2)
− I −1
− ∂I −1
A
∂MT
− I −1
−∂I −1
B
∂MT
((cid:126)ω1 + (cid:126)ω2)
B ((cid:126)ω1 + (cid:126)ω2)
(C.24)
(C.25)
(C.26)
∂ I B
∂(cid:126)θ2
(cid:16)(cid:16)
(cid:17)s(cid:17)
∂I −1
B
∂MT
∂2 (cid:126)ω2
∂(cid:126)ω1∂MT
+I −1
=
(cid:16)(cid:16) ∂I B
ω1I A −(cid:16)
∂MT
B
(cid:16)
I A(cid:126)ω1
I B((cid:126)ω1 + (cid:126)ω2)
(cid:17)s − ω1
(cid:16)
+ I −1
A
ω1
((cid:126)ω1 + (cid:126)ω2)
∂I −1
A
∂MT
+
(cid:16) − ω1I B − I B + ((cid:126)ω1 + (cid:126)ω2)
∂2 (cid:126)ω2
∂(cid:126)ω2∂MT
∂I −1
B
∂MT
=
(cid:17)
(cid:126)ω1
(cid:17)s − ω1I B − I B
(cid:17)
−(cid:16) ∂I A
(cid:17)
∂I B
∂MT
∂I A
∂MT
− ∂ I B
∂MT
∂MT
∂ I B
∂(cid:126)ω2
∂2 I B
∂(cid:126)ω2∂MT
−I −1
B ω1
∂I B
∂MT
− ∂ I B
∂MT
+ ((cid:126)ω1 + (cid:126)ω2)
40
(C.27)
(cid:17)s(cid:17)
(C.28)
Partials of A with respect to T jkl
Appendix C.2. Second Order Inertia Integral Partials of Dynamics Matrix
, where i represents body A or B and
j, k, l = 0 or 2 with only one index set to 2, are needed to compute the
sensitivity of the eigenvalues to the mass parameters of interest.
i
∂2(cid:126)r
= r
i
∂(cid:126)r∂T jkl
I A(cid:126)ω1 ω1 + I −1
A
−(cid:16) ∂I −1
A
∂T jkl
i
∂ (cid:126)MA
∂(cid:126)r
∂2 (cid:126)MA
∂(cid:126)r∂T jkl
(cid:126)MA + I −1
i
A
A
+ I −1
∂I −1
A
∂T jkl
i
(C.29)
(cid:17)s
∂ (cid:126)MA
∂T jkl
i
(cid:17)
(cid:16) ∂I −1
A
∂T jkl
i
∂I A
∂T jkl
− 1
m
i
(cid:126)ω1 ω1 +
∂3U
∂ (cid:126)R2∂T jkl
i
(cid:16) ∂I −1
A
∂T jkl
∂2(cid:126)r
∂(cid:126)θ2∂T jkl
i
= r
∂2(cid:126)r
∂(cid:126)ω1∂T jkl
i
m
(cid:17) − 1
(cid:16) ∂I A
(cid:17)
A
i
∂T jkl
∂3U
2∂T jkl
∂(cid:126)θ2
i
(cid:17)s
(cid:126)ω1
∂ (cid:126)MA
∂(cid:126)θ2
+ I −1
A
∂2 (cid:126)MA
∂(cid:126)θ2∂T jkl
i
(cid:17)s
(cid:16)
i
= r
(cid:16) ∂I −1
− ∂I −1
A
∂T jkl
i
i
I A(cid:126)ω1
A
∂T jkl
ω1I A − I −1
A ω1
+ I −1
∂I A
∂T jkl
i
∂2 (cid:126)ω1
∂(cid:126)r∂T jkl
i
=
∂I −1
A
∂T jkl
i
∂ (cid:126)MA
∂(cid:126)r
+ I −1
A
∂2 (cid:126)MA
∂(cid:126)r∂T jkl
i
∂2 (cid:126)ω1
∂(cid:126)θ2∂T jkl
i
=
∂I −1
A
∂T jkl
∂ (cid:126)MA
∂(cid:126)θ2
+ I −1
A
∂2 (cid:126)ω1
∂(cid:126)ω1∂T jkl
i
=
∂I −1
A
∂T jkl
− ∂I −1
A
∂T jkl
i
i
i
(cid:16)
(cid:17)s
I A(cid:126)ω1
+ I −1
A
ω1I A − I −1
A ω1
∂I A
∂T jkl
i
∂2 (cid:126)MA
∂(cid:126)θ2∂T jkl
i
(cid:16) ∂I A
(cid:126)ω1
∂T jkl
i
(cid:17)s
41
(C.30)
(C.31)
(C.32)
(C.33)
(C.34)
∂2 (cid:126)ω2
∂(cid:126)r∂T jkl
i
=
∂I −1
B
∂T jkl
i
∂ (cid:126)MB
∂(cid:126)r
+ I −1
B
(cid:16)
(cid:16)
∂2 (cid:126)ω2
∂(cid:126)θ2∂T jkl
i
=
∂I B
∂T jkl
− ∂I −1
B
∂T jkl
i
ω1
i
∂ (cid:126)MA
∂(cid:126)r
i
− ∂I −1
A
(cid:16)
∂T jkl
(cid:16)
+ I B
i
∂2 (cid:126)MB
∂(cid:126)r∂T jkl
∂I −1
∂(cid:126)θ2
(cid:17)
(cid:17)∂I B
ω1
B
∂(cid:126)θ2
i
A
− I −1
∂2 (cid:126)MA
∂(cid:126)r∂T jkl
∂2I −1
∂(cid:126)θ2∂T jkl
(cid:17)
(cid:17) ∂2I B
ω1
B
i
∂(cid:126)θ2∂T jkl
i
(cid:126)ω1 + (cid:126)ω2
(cid:126)ω1 + (cid:126)ω2
(C.35)
(C.36)
+
∂ (cid:126)MB
∂T jkl
∂2 (cid:126)MB
∂(cid:126)θ2∂T jkl
− ∂I −1
B
∂T jkl
i
i
i
B
(cid:126)ω1 + (cid:126)ω2
∂I −1
∂(cid:126)θ2
− ∂ I B
(cid:16)
∂T jkl
(cid:126)ω1 + (cid:126)ω2
i
+ (cid:126)MB
(cid:16)
(cid:17)∂ I B
(cid:126)ω1 + (cid:126)ω2
B
− I −1
B ω1
∂2I −1
∂(cid:126)θ2∂T jkl
(cid:17)∂I −1
(cid:16)
∂(cid:126)θ2
− I −1
B
i
B
− ∂I −1
A
(cid:16)
(cid:16)
∂T jkl
i
∂(cid:126)θ2
∂ (cid:126)MA
∂(cid:126)θ2
− I −1
(cid:17)(cid:17)s
I B
(cid:126)ω1 + (cid:126)ω2
+I −1
B
∂2 (cid:126)ω2
i
=
∂I −1
B
∂T jkl
− ∂I −1
(cid:16)
B
∂T jkl
i
I A(cid:126)ω1
∂(cid:126)ω1∂T jkl
i
− ∂I −1
A
∂T jkl
i
∂2 (cid:126)ω2
∂(cid:126)ω2∂T jkl
i
∂I B
∂T jkl
i
ω1I B − I −1
(cid:17)s − I −1
B ω1
(cid:16) ∂I A
A
(cid:126)ω1
∂T jkl
i
(cid:126)ω1 + (cid:126)ω2
∂I −1
B
∂T jkl
+
− I B
i
∂ (cid:126)MB
∂(cid:126)θ2
(cid:16)
(cid:17) ∂2 I B
(cid:126)ω1 + (cid:126)ω2
(cid:126)ω1 + (cid:126)ω2
∂(cid:126)θ2∂T jkl
i
(cid:17) ∂2I −1
B
∂(cid:126)θ2∂T jkl
i
i
A
B
∂2 (cid:126)MA
∂(cid:126)θ2∂T jkl
(cid:16) ∂I B
+ I −1
− ∂I −1
(cid:17)s
B
∂T jkl
I −1
∂T jkl
− ∂I −1
(cid:16)
B
∂T jkl
+
A
i
i
i
(cid:126)ω1 + (cid:126)ω2
(cid:16)
(cid:17)(cid:17)s
(C.37)
(cid:126)ω1 + (cid:126)ω2
i
∂T jkl
I B − I −1
B
∂ I B
∂T jkl
ω1I A + I −1
A ω1
i
∂I A
∂T jkl
i
∂ I B
∂T jkl
i
(C.38)
= − ∂I −1
(cid:16)
B
∂T jkl
− ∂I −1
B
∂T jkl
i
i
ω1I B − I −1
B ω1
(cid:17)∂ I B
∂I B
∂T jkl
− I −1
i
B
∂(cid:126)ω2
(cid:126)ω1 + (cid:126)ω2
I B − I −1
(cid:17) ∂2 I B
B
∂(cid:126)ω2∂T jkl
i
Appendix D. General Use Binary Asteroid Simulator
As a part of the analysis for this paper we developed a tool for dynamical
propagation of binary asteroids of arbitrary shape and expansion order using
42
Eq. 11-13 and the Hou mutual gravity potential described in Appendix A[11].
We have provided the software tool for free use at https://github.com/alex-
b-davis/gubas. The tool, referred to as the General Use Binary Asteroid
Simulator, is intended to provide the planetary science community with an
easily used, fast, and high fidelity simulation tool for the numerical inte-
gration of binary asteroid dynamics. It does not include the tools for the
fundamental frequency analysis performed in this paper.
The software was designed and implemented to be highly modular to en-
able a wide set of uses and allow for easy integration into larger tool sets. For
this reason the architecture was centered around a C++ executable wrapped
in a Python shell. The C++ executable performs the numerical integra-
tion and calculation of the inertia integrals while the Python wrapper pre-
processes user input from a configuration file to initialize the executable and
post-processes the results. This approach allows the user to easily modify
the Python shell script to fit their needs. In the standard architecture all
interactions are handles through the configuration file and a single command
line call to initialize the process. While the software and a detailed user guide
can be found by following the github link, Fig. B.11 shows a basic flowchart
of the software process.
Figure D.11: General Use Binary Asteroid Simulator Software Flowchart
43
References
[1] K. S. Noll, W. M. Grundy, E. I. Chiang, J.-L. Margot, S. D. Kern,
Binaries in the kuiper belt, arXiv preprint astro-ph/0703134.
[2] J.-L. Margot, M. Nolan, L. Benner, S. Ostro, R. Jurgens, J. Giorgini,
M. Slade, D. Campbell, Binary asteroids in the near-earth object pop-
ulation, Science 296 (5572) (2002) 1445 -- 1448.
[3] P. Pravec, P. Scheirich, P. Kusnir´ak, L. Sarounov´a, S. Mottola, G. Hahn,
P. Brown, G. Esquerdo, N. Kaiser, Z. Krzeminski, et al., Photometric
survey of binary near-earth asteroids, Icarus 181 (1) (2006) 63 -- 93.
[4] E. G. Fahnestock, D. J. Scheeres, Simulation of the full two rigid body
problem using polyhedral mutual potential and potential derivatives ap-
proach, Celestial Mechanics and Dynamical Astronomy 96 (3-4) (2006)
317 -- 339.
[5] S. P. Naidu, J.-L. Margot, Near-earth asteroid satellite spins under spin --
orbit coupling, The Astronomical Journal 149 (2) (2015) 80.
[6] A. J. Maciejewski, Reduction, relative equilibria and potential in the two
rigid bodies problem, Celestial Mechanics and Dynamical Astronomy
63 (1) (1995) 1 -- 28.
[7] P. Tricarico, Figure -- figure interaction between bodies having arbitrary
shapes and mass distributions: a power series expansion approach, Ce-
lestial Mechanics and Dynamical Astronomy 100 (4) (2008) 319 -- 330.
[8] D. Scheeres, Stability of the planar full 2-body problem, Celestial Me-
chanics and Dynamical Astronomy 104 (1-2) (2009) 103 -- 128.
[9] G. Bou´e, J. Laskar, Spin axis evolution of two interacting bodies, Icarus
201 (2) (2009) 750 -- 767.
[10] J. Ashenberg, Mutual gravitational potential and torque of solid bod-
ies via inertia integrals, Celestial Mechanics and Dynamical Astronomy
99 (2) (2007) 149 -- 159.
[11] X. Hou, D. J. Scheeres, X. Xin, Mutual potential between two rigid
bodies with arbitrary shapes and mass distributions, Celestial Mechanics
and Dynamical Astronomy (2016) 1 -- 27.
44
[12] Y. Takahashi, M. W. Busch, D. J. Scheeres, Spin state and moment
of inertia characterization of 4179 toutatis, The Astronomical Journal
146 (4) (2013) 95.
[13] A. I. Ermakov, M. T. Zuber, D. E. Smith, C. A. Raymond, G. Balmino,
R. R. Fu, B. A. Ivanov, Constraints on vestas interior structure using
gravity and shape models from the dawn mission, Icarus 240 (2014)
146 -- 160.
[14] D. Scheeres, S. Hesar, S. Tardivel, M. Hirabayashi, D. Farnocchia,
J. McMahon, S. Chesley, O. Barnouin, R. Binzel, W. Bottke, et al.,
The geophysical environment of bennu, Icarus 276 (2016) 116 -- 140.
[15] M. W. Buie, C. B. Olkin, W. J. Merline, K. J. Walsh, H. F. Levi-
son, B. Timerson, D. Herald, W. M. Owen Jr, H. B. Abramson, K. J.
Abramson, et al., Size and shape from stellar occultation observations
of the double jupiter trojan patroclus and menoetius, The Astronomical
Journal 149 (3) (2015) 113.
[16] D. Scheeres, Minimum energy configurations in the n-body problem and
the celestial mechanics of granular systems, Celestial Mechanics and
Dynamical Astronomy 113 (3) (2012) 291 -- 320.
[17] D. Scheeres, Relative equilibria for general gravity fields in the sphere-
restricted full 2-body problem, Celestial Mechanics and Dynamical As-
tronomy 94 (3) (2006) 317 -- 349.
[18] G. G´omez, Dynamics and Mission Design Near Libration Points. Vol. 1.
Fundamentals: the Case of Collinear Libration Points (World scientific
monograph series in mathematics; vol. 2), World Scientific, 2001.
[19] D. J. Scheeres, Orbital motion in strongly perturbed environments: ap-
plications to asteroid, comet and planetary satellite orbiters, Springer,
2016.
[20] F. Nimmo, O. Umurhan, C. M. Lisse, C. J. Bierson, T. R. Lauer, M. W.
Buie, H. B. Throop, J. A. Kammer, J. H. Roberts, W. B. McKinnon,
et al., Mean radius and shape of pluto and charon from new horizons
images, Icarus 287 (2017) 12 -- 29.
45
|
1907.00571 | 1 | 1907 | 2019-07-01T06:55:02 | New constraints on the dust and gas distribution in the LkCa 15 disk from ALMA | [
"astro-ph.EP"
] | We search a large parameter space of the LkCa 15's disk density profile to fit its observed radial intensity profile of $^{12}$CO (J = 3-2) obtained from ALMA. The best-fit model within the parameter space has a disk mass of 0.1 $M_{\odot}$ (using an abundance ratio of $^{12}$CO/H$_2$ $=$ 1.4 $\times 10^{-4}$ in mass), an inner cavity of 45 AU in radius, an outer edge at $\sim$ 600 AU, and a disk surface density profile follows a power-law of the form $\rho_r \propto r^{-4}$. For the disk density profiles that can lead to a small reduced $\chi^2$ of goodness-of-fit, we find that there is a clear linear correlation between the disk mass and the power-law index $\gamma$ in the equation of disk density profile. This suggests that the $^{12}$CO disk of LkCa 15 is optically thick and we can fit its $^{12}$CO radial intensity profile using either a lower disk mass with a smaller $\gamma$ or a higher disk mass with a bigger $\gamma$. By comparing the $^{12}$CO channel maps of the best-fit model with disk models with higher or lower masses, we find that a disk mass of $\sim$ 0.1 $M_{\odot}$ can best reproduce the observed morphology of the $^{12}$CO channel maps. The dust continuum map at 0.87 mm of the LkCa 15 disk shows an inner cavity of the similar size of the best-fit gas model, but its out edge is at $\sim$ 200 AU, much smaller than the fitted gas disk. Such a discrepancy between the outer edges of the gas and dust disks is consistent with dust drifting and trapping models. | astro-ph.EP | astro-ph | Draft version July 2, 2019
Typeset using LATEX default style in AASTeX62
9
1
0
2
l
u
J
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
7
5
0
0
.
7
0
9
1
:
v
i
X
r
a
New constraints on the dust and gas distribution in the LkCa 15 disk from ALMA
Sheng Jin,1 Andrea Isella,2 Pinghui Huang,1, 2, 3, 4 Shengtai Li,3 Hui Li,3 and Jianghui Ji1
1CAS Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China
2Department of Physics and Astronomy, Rice University, 6100 Main St., Houston, TX 77005, USA
3Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA
4University of Chinese Academy of Sciences, Beijing 100049, Peoples Republic of China
(Received March 25, 2019)
Submitted to ApJ
ABSTRACT
We search a large parameter space of the LkCa 15's disk density profile to fit its observed radial
intensity profile of 12CO (J = 3-2) obtained from ALMA. The best-fit model within the parameter
space has a disk mass of 0.1 M(cid:12) (using an abundance ratio of 12CO/H2 = 1.4 ×10−4 in mass), an
inner cavity of 45 AU in radius, an outer edge at ∼ 600 AU, and a disk surface density profile follows
a power-law of the form ρr ∝ r−4. For the disk density profiles that can lead to a small reduced χ2 of
goodness-of-fit, we find that there is a clear linear correlation between the disk mass and the power-law
index γ in the equation of disk density profile. This suggests that the 12CO disk of LkCa 15 is optically
thick and we can fit its 12CO radial intensity profile using either a lower disk mass with a smaller γ or
a higher disk mass with a bigger γ. By comparing the 12CO channel maps of the best-fit model with
disk models with higher or lower masses, we find that a disk mass of ∼ 0.1 M(cid:12) can best reproduce the
observed morphology of the 12CO channel maps. The dust continuum map at 0.87 mm of the LkCa
15 disk shows an inner cavity of the similar size of the best-fit gas model, but its out edge is at ∼ 200
AU, much smaller than the fitted gas disk. Such a discrepancy between the outer edges of the gas and
dust disks is consistent with dust drifting and trapping models.
Keywords: stars: individual (LkCa 15) -- protoplanetary disks -- radiative transfer -- submillimeter:
planetary systems
1. INTRODUCTION
Recently, a large number of protoplanetary disks were spatially resolved by the Atacama Large Millime-
ter/submillimeter Array (ALMA) and the Next Generation Very Large Array (ngVLA)(ALMA Partnership et al.
2015; Fedele et al. 2017; Andrews et al. 2018; Isella et al. 2018; Liu et al. 2018). These high-resolution observations
of protoplanetary disks in dust continuum and molecular line emissions provide us the morphology of dust and gas
distributions in a wide variety of protoplanetary disks. Such information places fundamental constraints for the theo-
retical studies on dust properties and dust-gas interaction, which is an important building block of planet formation
theory (Dong et al. 2015; Isella et al. 2016; Jin et al. 2016; Liu et al. 2018; Dong et al. 2018; Huang et al. 2018; Ricci
et al. 2018; Huang et al. 2019).
LkCa 15 is a 2-5 Myr old K5 star with L(cid:63) ∼ 0.74 L(cid:12) and M(cid:63) ∼ 1.0 M(cid:12) (Kenyon & Hartmann 1995; Simon et al.
2000). It is located in the Taurus-Auriga star-forming region at a distance of 140 pc from the Earth(van den Ancker
et al. 1998). LkCa 15 is an interesting target due to its partially dust-depleted disk (Pi´etu et al. 2006; Espaillat et al.
2007; Thalmann et al. 2010; Andrews et al. 2011; Isella et al. 2012, 2014; Thalmann et al. 2014) and the probability of
harboring a planet candidate inside its inner cavity (Kraus & Ireland 2012; Sallum et al. 2015; Thalmann et al. 2016;
Corresponding author: Sheng Jin
[email protected]
2
Jin et al.
Currie et al. 2019). The depleted inner region shown in the dust continuum image is about 50 AU in radius (Pi´etu et al.
2006; Andrews et al. 2011; Isella et al. 2012, 2014), and it has a mass accretion rate of about 10−9 M(cid:12)yr−1(Hartmann
et al. 1998). The outer edge of the dust disk inferred from continuum emission is at ∼ 150 AU(Pi´etu et al. 2007; Isella
et al. 2012). The dust mass estimated from 1.3 mm continuum observation is about 5 × 10−4 M(cid:12)(Isella et al. 2012).
Recent images of scattered light suggest a warped inner disk component inside the inner gap, providing a clear picture
details of the inner gap region of the LkCa 15 disk (Thalmann et al. 2015, 2016; Oh et al. 2016).
As a young star, LkCa 15 is a luminous source of X-ray and EUV emission(Skinner & Gudel 2013), indicating that
the LkCa 15 disk could still be undergoing the active evolution phase of disk physics and chemistry. Consequently,
the protoplanetary disk of LkCa 15 has been found to be especially chemically rich and has been detected in several
molecular transitions(Thi et al. 2004; Pi´etu et al. 2007; Chapillon et al. 2008; Oberg et al. 2010; Punzi et al. 2015).
Molecular line emission shows that the gas disk around LkCa 15 is in size of ∼ 900 AU(Pi´etu et al. 2007; Isella et al.
2012). It is highly optically thick in the emission of 12CO, but optically thin in 13CO emission(Punzi et al. 2015; van
der Marel et al. 2015). The discrepancy between the outer disk radii shown in the dust continuum and molecular line
emissions (∼ 150 AU versus 900 AU) suggests that the mm-size dust is depleted in the outer part of the LkCa 15 disk,
which is in consistence with dust drifting and trapping models(Birnstiel et al. 2010; Pinilla et al. 2012).
Being a protoplanetary disk that may have a young planet candidate, the LkCa 15 system serves as a unique
laboratory to study the dust evolution models under planet-disk interaction(Zhu et al. 2011; Pinilla et al. 2012). The
distribution of dust and gas in the LkCa 15 disk provides an important constraint on theoretic models. Here, we fit
the gas and dust surface density profiles based on the high-resolution dust continuum image and CO 3-2 line maps
obtained from the ALMA. We study how the goodness-of-fit changes along with different key parameters used in
the disk density profile, which is related with different physics. Our disk model and the parameter grid is described
In Section 4, we show the best-fit model with respect to the observed 12CO image. We discuss the
in Section 3.
dependence of parameters in Section 5.
2. OBSERVATION
The LkCa 15 was observed at August 17 and August 29, 2014 with ALMA Band 7 (345 GHz, 880 µm) in ALMA
program 2012.1.00870.S (PI P´erez, L. M.). The 12CO, 13CO and C18O 3-2 line maps and 0.87 mm dust continuum
image were taken with two different tunings. The 12CO 32 data were obtained with a spectral window centered on
345.796 GHz, with 1885 channels of 122.06 kHz (0.106 km s−1) channel width. The angular resolution of the 12CO
image is 0.36.(cid:48)(cid:48)×0.23.(cid:48)(cid:48) (50 AU x 32 AU). The 13CO 32 data were centered on 330.588 GHz, with 1967 channels of
122.06 kHz (0.11 km s−1) channel width. The angular resolution of the 13CO image is 0.28.(cid:48)(cid:48)×0.21.(cid:48)(cid:48) (40 AU x 30 AU).
The C18O 32 data were centered on 329.331 GHz, with 1967 channels of 122.06 kHz (0.111 km s−1) channel width.
The angular resolution of the C18O image is 0.30.(cid:48)(cid:48)×0.23.(cid:48)(cid:48) (41 AU x 32 AU). The dust continuum observation consists
of four spectral windows(334.01∼335.99 GHz, 339.02∼341.00 GHz, 341.02∼343.00 GHz, 346.01∼347.99 GHz). Each
spectral window has 64 channels of 31248.22kHz (3527.09 km s−1) channel width. The angular resolution of the dust
continuum image is 0.23.(cid:48)(cid:48)×0.17.(cid:48)(cid:48) (32 AU x 23 AU).
This is the first comprehensive observation of the CO 3-2 line emission of 12CO, 13CO and C18O in the LkCa 15
disk. Previous observations revealed the line emission of various molecules in the LkCa 15 disk, including the 12CO
6-5(van der Marel et al. 2015), CO 2-1 and HCO+(Qi et al. 2003; Pi´etu et al. 2007; Punzi et al. 2015), S-bearing
molecules(Dutrey et al. 2011), and ethynyl radical (CCH)(Henning et al. 2010). The 12CO 2-1 line emission shows
that the LkCa 15 disk is highly optically thick in 12CO(Punzi et al. 2015), and the 12CO 6-5 emission show that gas
is still present inside the observed dust cavity(van der Marel et al. 2015). The dust continuum image has also been
obtained by several former millimeter observations(Pi´etu et al. 2006; Espaillat et al. 2007; Isella et al. 2012, 2014), and
all these observations show an inner dust cavity of ∼ 40-50 AU in size.
2.1. Dust continuum and line maps
Figure 1 shows the dust continuum and the 12CO zero-moment maps of the LkCa 15 system. We will extract an
azimuthal averaged flux from these zero-moment maps along the radial direction, and use the derived 12CO radial
intensity profile as a reference to judge the goodness-of-fit of our disk models. The dust continuum map shows an
inner hole of ∼ 65 AU in radius, and the full width at half maximum of this dust cavity is at ∼ 40 AU. This is in
agreement with previous millimeter observations(Andrews et al. 2011; Isella et al. 2012, 2014). Such a large inner cavity
disappears in the zero-moment map of 12CO. For the optically thin 13CO and C18O emissions, the inner disk show an
LkCa 15 manuscript
3
obvious decrease in the azimuthal averaged flux. The zero-moment maps of 12CO, 13CO and C18O are consistent with
previous findings that the LkCa 15 disk is optically thick in 12CO emission, while optically thin in 13CO and C18O.
Figure 2 shows the observed channel maps of 12CO, 13CO and C18O of the LkCa 15 disk. The channel maps of
12CO clearly show the near and far halves of a double cone structure, which is the feature resulted from a circular
Keplerian rotational disk(Rosenfeld et al. 2013). However, we do not observe such a feature in the channel maps of
the optically thin 13CO and C18O due to lower masses of these two isotopes.
3. MODELING
3.1. surface density profile and parameter grid
Our primary goal here is to find a gas disk surface density profile that can best reproduce the observed zero-moment
12CO map. First, we create a parameter space by parameterizing the surface density equation of an analytical disk
model. Then, we search the parameter space to obtain the best-fit model that has the least reduced χ2 of the radial
intensity profile of 12CO emission.
The observed radial intensity profiles of CO isotopes and dust continuum emission in the LkCa 15 disk were obtained
by extracting the azimuthal averaged intensity of the zero-moment maps shown in Figure 1. The radial intensity profiles
of CO are extended to ∼ 600 AU, while the dust continuum emission shows a ring-shape structure and ends at ∼ 200
AU. The dust continuum and the CO line emissions cannot be fitted using a single surface density profile. Thus we
use two separate equations to describe the surface density profiles of dust and gas in the LkCa 15 disk.
We employ parameterized analytical disk surface density profiles to model the gas and dust in the LkCa 15 disk.
For the gas, we adopt a surface density profile that is described by
Σg(r) = Σ0
arctan(r/RCarctan)γarctan
,
(1)
(cid:16) r
(cid:17)−γ
RC
which is a simple power-law density profile combined with an inner cavity described by an arctangent function. We
turn off the exponential decay term that are typically used to describe the surface density of protoplanetary disks
(e.g., Andrews et al. 2009) because the intensity of CO line emissions decrease slowly at larger radii. Moreover, we
fix the RC at 12500 AU to slow down the decrease of the gas density at the outer part of the disk. There are four
free parameters in Equation 1: Σ0 that determines the disk mass, γ the power-law decay of the surface density in the
radial direction, RCarctan the size of the inner cavity, and γarctan the slope of the junction region between the inner
cavity and the outer disk. We set up a four-dimensional parameter space of these four free parameters. The parameter
grids at each dimension are listed in Table 1.
The dust surface density profile is described by
Σd(r) = Σ0
exp[−(r/RC)2−γ] arctan(r/RCarctan)γarctan
(2)
(cid:16) r
(cid:17)−γ
RC
Compared with the surface density profile of the gas disk, there is an exponential decay term in the dust density profile
to simulate the disappearance of the dust intensity at larger radii. Since we have to subtract the dust continuum
emission in generating the 12CO images for all the models in the parameter space listed by Table 1, we adopt a
fixed dust density profile of RC = 66 AU, γ = −0.15, RCarctan = 66 AU, γarctan = 5.35, and a total dust mass of
Mdust = 9.8 × 10−5M(cid:12). These values are determined by fitting the azimuthal averaged radial intensity profile of the
observed dust continuum map, and they are used for all the 4096 runs in the parameter space.
3.2. physical model
The first step to derive the 12CO intensity from a specific gas surface density profile is to calculate the three-
dimensional disk temperature structure. For each gas surface density profile in the parameter space, We use an iterative
approach to obtain a self-consistent three-dimensional disk temperature structure. We assume the disk temperature
structure is controlled by micron-size dust particles that are well coupled with gas. Since the vertical distribution
of micron-size dust is in turn determined by the disk temperature structure, this is a circular dependency problem.
To solve this problem, first we generate an initial three-dimensional micron-size dust distribution, and calculate an
initial disk temperature based on this dust distribution. Using the calculated disk temperature, we produce a new
micron-size dust distribution by solving the differential equation of hydrostatic equilibrium:
− ∂ ln ρgas
∂z
=
∂ ln Tgas
∂z
+
1
2
cs
(cid:20) GM∗z
(cid:21)
(r2 + z2)3/2
,
(3)
4
Jin et al.
2 = kBTgas/µmh is the sound speed. We then run Monte Carlo radiation transfer simulation to calculate a new
where cs
disk temperature structure using the updated dust distribution. We repeat this process until the three-dimensional
disk temperature structure used in generating a dust distribution and the disk temperature structure calculated by
the same dust distribution converge (the difference between two temperature structures is within 3% everywhere).
The disk temperature structure and the dust continuum and CO line maps are calculated using the Monte Carlo
radiative transfer code RADMC-3D(Dullemond 2012). We assume the host star of LkCa 15 is a black-body radiator
with an effective temperature of 4350 K. The surface density of micron-size dust in the calculation of the disk tem-
perature profile is obtained by a dust-to-gas ratio of 0.001. Compared to the Milky Way average ratio of 0.01 (Bohlin
et al. 1978), by setting such a small dust-to-gas ratio we assume that substantial grain growth has occurred in the
LkCa 15 disk and the micron-size dust is about 10% in mass of the total dust. Submillimeter and dust continuum
observations show that dust grain growth to mm-size particles is completed within less than 1 Myr for the majority of
circumstellar disks(Rodmann et al. 2006; Draine 2006; Ricci et al. 2010a,b, 2011; Ubach et al. 2012), and simulation
of dust evolution also shows mm-size dust can form at the age of ∼ 1 Myr at high density region in protoplanetary
disk(Ormel et al. 2009). As aforementioned, the LkCa 15 disk is about 2-5 Myr in age(Kenyon & Hartmann 1995;
Simon et al. 2000), thus a large fraction of dust can appear to be in large sizes at this stage. The dust opacities used
in this work were calculated similar to that of Isella et al. (2009). We assume the dust grains are compact spheres
made of astronomical silicates (Weingartner & Draine 2001), organic carbonates (Zubko et al. 1996), and water ice,
with fractional abundances described in Pollack et al. (1994). Single grain opacities were averaged on a grain size
distribution to obtain the mean opacity. We adopt a typical MNR power-law size distribution (Mathis et al. 1977),
n(a) ∝ a−3.5, between a minimum grain size of 5× 10−6 mm and a maximum grain size of 1× 10−2 mm. The resulting
dust opacity at the wavelength of 1 µm is 5.2 × 103 cm2 g−1.
After we have obtained a self-consistent three-dimensional disk temperature structure, we calculate the dust con-
tinuum and 12CO line emission. We interpolate the surface density profile of mm-size dust given by Equation 2 on a
three-dimensional spherical grid with a scale height profile of hmm−dust(r) = 0.1 × 1.0AU × (r/20AU)1.25, where 0.1 is
a parameter that accounts for the settling of 0.15 millimeter size dust towards the mid-plane. This results in a scale
height of ∼ 0.75 AU at 100 AU for relatively large dust particles. Such a scale height is consistent with the findings
by Pinte et al. (2015). The dust opacity adopted for mm-size dust is calculated using the same model of the opacity
for micron-size dust, the only difference is that here the maximum grain size is 1 mm. The resulting dust opacity at
the wavelength of 1 mm is 13.2 cm2 g−1. We convolve the dust continuum map in each model with the PSF of the
ALMA observation and extract an azimuthal averaged radial intensity profile. Then we calculate the reduced χ2 for
dust continuum emission between the radial intensity profiles of the model and the ALMA observation. For different
models in our parameter space, the goodness-of-fit of the dust continuum emission changes due to the variation of disk
temperature. But we find that the influence of dust continuum map on the zero-moment 12CO map is limited, because
the dust continuum image has a ring-like structure that located at the optically thick region of the 12CO emission.
To calculate the line emission of 12CO, we assume the abundance ratio of 12CO/H2 to be 1.4 ×10−4 in mass, and this
value is consistent with the canonical ratio of 104 for the disk initial conditions(Lacy et al. 1994; France et al. 2014).
Then we create a three-dimensional density structure of 12CO based on the gas surface density profile of each run by
solving the Equation 3. We include the freeze-out effect of CO by setting the number density of 12CO to zero at the
region where temperature is below 20 K. Photodissociation by stellar UV and/or X-ray radiation is another important
factor as it can destroy the CO in the surface layers of the disk(Visser et al. 2009). We follow the same procedure as in
that of Qi et al. (2011) and Rosenfeld et al. (2013), in which we calculate a photodissociation boundary by vertically
integrating the H nuclei density to a threshold density of 2.0 ×1020. This is a mild threshold compared to the value of
5.0 ×1020 used in Rosenfeld et al. (2013). We find that it is hard to fit the 12CO intensity beyond ∼ 400 AU using a
larger photodissociation threshold like 5.0 ×1020 in H nuclei, because with such a strong photodissociation rate there
are little 12CO in the outer disk. The photodissociation threshold turns out to be an alternative free parameter in the
fitting of the 12CO radial intensity profile, as it is a critical parameter for the 12CO intensity in the outer disk. In this
work, we fix the photodissociation threshold of 2.0 ×1020 for all the 4096 runs.
We first calculate the self-consistent disk temperature. Then we calculate the dust continuum map and the channel
maps of 12CO, and convolve these images with the corresponding ALMA PSFs. Afterwards we subtract the dust
continuum map from each channel map, and integrate all the channel maps to generate a zero-moment map of 12CO.
Finally, we subtract an azimuthal averaged radial intensity profile from the zero-moment 12CO map and calculate the
corresponding reduced χ2 of 12CO emission.
LkCa 15 manuscript
5
In addition, once we obtain the best-fit parameter sets for the 12CO emission, we tentatively adjust the abundance
ratio of 12CO/13CO to fit the radial intensity profile of 13CO, using the same disk temperature profile given by the
best-fit 12CO model. The observed 13CO radial intensity profile shows a peak at ∼ 45 AU, although there is no such
a peak in the radial intensity profile of 12CO (the radial intensity profiles of 12CO and 13CO can be found in Section
4). Hence, we have to adopt different abundance ratios of 12CO/13CO for the disk region inside or outside of 45 AU.
Although the fitting process of 13CO is not a self-consistent approach, the result can at least be an estimation of the
number density of 13CO. We do not model the C18O disk because the signal-to-noise ratio is low and no flux are
detected outside of 200 AU.
4. THE BEST-FIT MODEL
4.1. 12CO intensity profile
The best-fit model in our parameter space has the following parameters: Mdisk = 0.1M(cid:12), γ = 4.0, RCarctan = 45
AU, and γarctan = 10. Although the fitted disk mass is relatively large, it agrees with previous findings(Isella et al.
2012; Huang et al. 2017). Figure 3 shows the gas surface density profile of this best-fit model, which is related to
a power-law disk without an exponential decay term usually used in the description of protoplanetary disks (e.g.,
Andrews et al. 2009). Figure 4 gives the self-consistent temperature structure as calculated by our iterative approach
that was described in Section 3.2. It shows a typical two-layer vertical structure of passive irradiated circumstellar
disks (Dullemond et al. 2002). In the surface layer, the temperature decreases from ∼ 51 K at 100 AU to ∼ 29 K at
500 AU. In the mid-plane, the temperature decreases from ∼ 15 K at 100 AU to ∼ 11 K at 500 AU.
The top panel of Figure 5 compares the azimuthal averaged radial intensity profiles extracted from the zero-moment
12CO maps of our best-fit model with the ALMA observation. The best-fit model results in a reduced χ2 of 2.51. Its
gas surface density profile has an inner cavity of 45 AU in size. However, we do not see such a cavity in the resulting
radial intensity profile, since most of the inner cavity is optically thick in 12CO emission. In fact, for a majority part
of the gas density profiles in our parameter space, the inner disk region is optically thick in 12CO emission and its
intensity only depends on the calculated disk temperature. Although the very inner part of the disk cavity can become
optically thin due to the arctangent function used in the surface density equation, this inner optically thin region did
not show up in the radial intensity profile because of the large PSF of the ALMA observation, which is of ∼ 50 AU in
major axis and ∼ 32 AU in minor axis. The observed 12CO intensity decreases slowly in the outer optically thin part
of the disk. We find that it is hard to fit the slope of the 12CO radial intensity profile in the entire disk using a simple
power-law surface density. Our best-fit model has a large γ of 4.0. It fits the inner ∼ 400 AU very well, but beyond
∼ 400 AU it shows lower intensity compared to the ALMA observation. Note that the photodissociation threshold
can be another free parameter in our model that can affect the intensity in the outer part of the disk. In order to
slow down the decrease of the 12CO intensity beyond ∼ 400 AU, we set a photodissociation threshold of 2.0 ×1020 in
H nuclei to keep more 12CO in the outer disk, which is a mild threshold compared to the value of 5.0 ×1020 used in
Rosenfeld et al. (2013). Different photodissociation thresholds will result in different best-fit parameter sets.
In Figure 6 and 7 we show the dust continuum map, the 12CO zero-moment map, and the 12CO channel maps of our
best-fit model. Rather than interpolating the Fourier transformation of our model images to the actual observation
dataset and cleaning the dataset to get exactly the same PSF of ALMA, we simply convolve a Gaussian function of
the same sizes of major and minor axes with the ALMA observation to speed up our fitting process. Thus, compared
to Figure 1 and 2, our model images are more smooth than the ALMA observation. For the calculation of reduced χ2,
we only use the azimuthal averaged intensity profile. We notice that the channel maps provide important constraints
on the goodness of fitting. In the 12CO channel maps shown in Figure 7, we can clearly see the near and far halves
of a double cone structure of a Keplerian disk, and the relative angle and magnitude of the near and far halves share
similarities with the observed structure shown in Figure 2. These similarities between the channel maps of our best-fit
model and the observed channel maps suggest that the mass of LkCa 15 disk is around 0.1 M(cid:12) under the assumption
that the abundance ratio of 12CO/H2 is ∼ 1.4 ×10−4 in mass. The fitting of the outer disk is not good as the disk size
is smaller in the channel maps shown in Figure 7, while the observed channel maps in Figure 2 exhibits more extended
images. This difficulty is due to the simple power-law surface density profile and the identical radial photodissociation
threshold used in our model. We will investigate the effect of different disk masses in Section 5.2. Note that the fitted
disk mass depends on the abundance ratio of 12CO/H2 in our model. If we adopt a larger abundance ratio of 12CO/H2
of ∼ 3.0 ×10−4 in mass, the fitted mass of the LkCa 15 disk should be around 0.05 M(cid:12).
6
Jin et al.
The bottom panel of Figure 5 compares the azimuthal averaged radial intensity profile of the dust continuum emission
of our best-fit model with the ALMA observation. The observation bias for dust continuum image is small, as shown
by the small error-bar in the observed radial intensity profile. As a result, the radial intensity profile extracted from
the dust continuum image of our best-fit model has a large reduced χ2 of 255. But the two intensity profiles of the
best-fit model and the ALMA observation generally match. The χ2 of dust continuum emission can be largely reduced
by further fine tuning the dust surface density profile.
4.2. 13CO image
Based on the surface density profile and the temperature structure of the best-fit model of 12CO, we manually adjust
the abundance ratio of 12CO/13CO to fit the observed 13CO intensity. This actually exhibits the 13CO emission at the
temperature profile given by the gas density profile obtained by the fitting of 12CO, and it is not a self-consistent way
as compared to the fitting of 12CO intensity. We adopt this simplified approach only to estimate the mass of 13CO
needed to reproduce the observed 13CO image based on the disk properties of the fitted 12CO disk.
We use a power-law function to describe the radius-dependent abundance ratios of 12CO/13CO:
(cid:16) r
(cid:17)α
η(r) = n0
45AU
,
(4)
where η(r) is the abundance ratio at r, and n0 is the abundance ratio at 45 AU. We separately fit the α for the inner
and outer disk regions by manually adjusting n0 and α. We obtained a n0 of 6360, α = -4.0 for the disk region inside
of 45 AU, and α = 1.7 outside of 45 AU. This leads to 12CO/13CO = 1, 248, and 6360 at 5, 20, and 45 AU, and
12CO/13CO = 1640, 500, 155 and 78 at 100, 200, 400 and 600 AU. Since the γ in the best-fit 12CO disk is 4, we can
infer that the number density profile of 13CO is of the form r−2.3 outside of 45 AU. This could partly explain why
it is difficult to fit the 12CO intensity in the outer disk using a surface density profile that is of the form r−4, given
that the number density profile fitted from the optically thin 13CO is of r−2.3. But if we use a surface density profile
that is of the form r−2.3, the fitted disk mass will be around 10−3 to 10−2 M(cid:12) (see Section 5.1 for details). Such a
low mass can not reproduce the morphology of the observed 12CO channel maps as shown in Section 5.2. Thus it is
difficult to fit the 12CO intensity in the inner and outer disk regions using a simple power-law radial density profile.
Figure 5 shows the radial intensity profile of 13CO compared with the ALMA observation. The reduced χ2 of 13CO is
derived to be 0.33.
The zero-moment and channel maps of 13CO obtained from the aforementioned fitted abundance ratios are shown
in Figure 6 and 7. The fitted zero-moment map of 13CO has an inner cavity, which means the disk becomes optically
thin at the wavelengths of 13CO line emissions. The observed zero-moment map of 13CO does not show a distinct
inner cavity, but we can see an obvious decrease of the intensity in the inner disk region. For the channel maps, since
the mass of 13CO is much lower than the 12CO, we do not see the near and far halves of a double cone structure as
shown in the channel maps of 12CO. The morphology of the 13CO channel maps agrees well with the observation.
5. A PARAMETER STUDY
5.1. dependence of parameters
We aim to investigate how the fitting of the 12CO intensity changes with different parameters. For this, we plot
in Figure 8 the heat maps of the reduced χ2 of the models that have at least two equal parameters with our best-fit
model. For example, The top left panel shows the 64 models that have the same RCarctan and γarctan as the best-fit
model (marked with a green circle). We see that for the models that have relatively lower reduced χ2, where they
show a clear linear correlation between the Mdisk and γ. This means that we can find a reasonable fit of the observed
12CO intensity using either a low disk mass with a small γ or a high disk mass with a large γ. This relation can be
explained as follows. For most of our 4096 models, the inner part of the disk (inside of ∼ 300 AU) is optically thick for
the line emission of 12CO. As a result, the disk temperature profile determines the observed intensity in the inner part
of the disk. Since all 64 runs in this panel have the same RCarctan and γarctan, they will have similar disk temperature
structures in the inner part of the disk. Thus the key factor to obtain a good fit for these 64 runs is to derive the
right intensity in the outer disk region. This means that in the case of a high mass disk, we only need to reduce the
mass at the outer disk region to obtain low intensity there to fit the observed 12CO emission. According to Equation
1, to achieve this point, we can set a large γ to decrease the disk surface density quickly at large radii. In Figure 5 we
also plot the intensity profiles of two disks that has a higher disk mass of of 0.33 M(cid:12) or a lower disk mass of 0.033
LkCa 15 manuscript
7
M(cid:12), as compared to the best-fit model. The intensity profiles of these three models in the inner 100 AU overlap each
other. The high-mass model result in a reduced χ2 of 14.6, and the low-mass model result in a reduced χ2 of 6.7. The
difference between three models is the goodness-of-fit at the outer disk region. We will show in Section 5.2 that the
goodness of these three models can also be clearly observed in the morphology of the 12CO channel maps.
The top right panel of Figure 8 shows the 64 models that have the same γ and γarctan with our best-fit model.
This panel shows that for the disks that have a small inner cavity (with size < 10 AU), they cannot reproduce the
observed 12CO intensity. Only models that have an inner cavity of ∼ 30-60 AU can possibly obtain a small reduced χ2.
Furthermore, we can see in the case that the γ, γarctan and Mdisk are fixed, the goodness-of-fit shows weak correlation
to the size of the inner cavity (RCarctan). On the other hand, when the γ, γarctan and RCarctan are fixed, the goodness
of fitting shows weak correlation to the disk mass, since the 12CO emission is optically thick.
The middle left panel of Figure 8 shows the 64 models that have the same Mdisk and γarctan with the best-fit model.
The row with γ = 4 shows again that the size of the inner cavity plays a less important role in the goodness of fitting.
The essential part in the fitting of the 12CO intensity is to find a combination of γ and Mdisk that can fit the outer
part of the disk. If this goal is achieved, then we can obtain a model that fit the observation, regardless small changes
in γarctan and RCarctan.
The middle right panel of Figure 8 shows how the reduced χ2 depends on γarctan and RCarctan, i.e., the size of the
inner cavity and the slope of the connection region between the inner cavity and outer disk. It confirms that these
two parameters have weak effect on the goodness-of-fit of the 12CO intensity. However, there should be a large inner
cavity. As it shows that an inner cavity of size < 20 AU do not reproduce the observed 12CO intensity.
The bottom two panels of Figure 8 shows how the reduced χ2 changes in the γ versus γarctan space and the Mdisk
versus γarctan space. The left panel shows that if the Mdisk is fixed, then the γ determines the goodness-of-fit. On the
contrary, the right panel shows that if the γ is fixed, the Mdisk determines the goodness-of-fit. For the γarctan, here we
see again that it has a very limited effect on the goodness-of-fit. Therefore, we may conclude that the most important
part in obtaining a good fit is to find a combination of Mdisk and γ.
5.2. Constraint on the disk mass
The best-fit model in our parameter space has a disk mass of 0.1 M(cid:12), a γ of 4.0, and an inner cavity of 45 AU.
We have seen in Section 5.1 that the goodness-of-fit is affected by different combination of parameters, as the top left
panel in Figure 8 shows that by adjusting the γ, we can obtain reasonable fits with higher or lower disk masses. Here
we investigate how much our model constrains the mass of the LkCa 15 disk. We choose three models from the top
left panel in Figure 8, the best-fit model, a high-mass disk model of 0.33 M(cid:12) and a low-mass model of 0.033 M(cid:12) (these
three models are marked with a green, brown and blue circle respectively). The radial intensity profiles of three models
shown in Figure 5, and the high-mass or low-mass model either over-produce or under-produce the intensity out of ∼
100 AU. Since the radial intensity profile is extracted from the zero-moment map that is actually a degenerated image
obtained by combining all the channel maps in an actual observation, we expect the difference of these models can
also be observed in the channel maps.
In Figure 9 we compare the three models' channel maps with the ALMA observation, where all the channel maps are
plotted in the same colorbar. The most apparent difference between three models is the size of the channel maps. Since
we have seen in Figure 5 that the high-mass and low-mass model result in either higher or lower intensity in the outer
part of the disk, they either show larger or smaller size in the channel maps compared with the best-fit model. There is
another feature that can be used as an effective criteria for the goodness-of-fit of the disk mass, i.e., the morphology of
the near and far halves of a double cone structure of a Keplerian disk. For example, in the channel map at the velocity
of ± 1.05km/s, the high-mass model shows a too much intense far halve in the double cone structure compared to
the observation, while the low-mass model shows a much smaller double cone structure since it under-produce the
intensity at the out disk region. In the channel maps at the velocity of ± 0.63 and ± 0.42km/s, the two short wings
of the double cone structure in the high-mass model are too strong compared to the observation. Thus, the inability
of the high-mass and low-mass model in fitting the intensity in the outer disk region also exhibits in the morphology
of the double cone structure in the channel maps. They show a double cone structure that is either too strong or too
weak compared to the observed channel maps.
The mm-size dust surface density profile that can reproduce the observed dust continuum map is significantly different
compared to the fitted gas surface density profile, and this is a reliable result because of the small observational bias
of the dust continuum emission. The mm-size dust density profile has a peak at ∼ 65 AU, indicating a pressure bump
Jin et al.
8
exists at the same location in the gas disk. The full width at half maximum of this dust cavity is at ∼ 40 AU, similar
as the inner cavity of the fitted gas disk. The mm-size dust disk has an outer edge at ∼ 200 AU, which is much
smaller compared to the fitted gas disk of ∼ 600 AU in size. Such a discrepancy between the gas and dust disks
provides an important constraint of the dust drifting models in the LkCa 15 system. Our mm-size dust disk is of ∼
1.0 ×10−4 M(cid:12) in mass. According to our dust opacity model, the opacity at mm wavelength is mainly contributed
by dust of ∼ 0.15 mm is size. This suggests a dust-to-gas ratio of ∼ 0.001 for dust of ∼ 0.15 mm in size. But this is
a weak constraint because we adopt a uniform opacity model for dusts at different distances. In reality, the species
and the size distribution of dust should change at different radii, and this will affect the calculated dust opacity and
consequently the fitted dust masses at different radii.
6. CONCLUSIONS
In this work, we analyze the dust continuum and 12CO 3-2 line emission maps of the LkCa 15 disk that are obtained
from ALMA observation. We parameterize an analytical surface density profile of the LkCa 15 disk and search
through the parameter space to find the best-fit gas surface density model that leads to the least reduced χ2 of the
12CO intensity. Our key findings are summarized as follows:
1. The best-fit model of the gas disk based on 12CO 3-2 line emission is a disk of 0.1 M(cid:12) in mass. The gas disk has
an inner cavity of 45 AU in size, and its outer edge is at ∼ 600 AU. The surface density profile of this best-fit model
follows a power-law of the form ρr ∝ r−4. But such a steep power-law density profile results in lower 12CO intensity
at outer part of the disk beyond ∼ 400 AU compared to the observation.
2, The dust continuum map can be reproduced by a dust disk which has an inner cavity of ∼ 65 AU in size, and the
full width at half maximum of this cavity is ∼ 40 AU. The size of the dust cavity is similar to the size of the fitted
gas cavity. Unlike the gas disk, the mm-size dust disk ends at ∼ 200 AU. The discrepancy between the outer edges of
the gas and dust disks can be used to study the dust drifting models in the LkCa 15 disk.
3, The heat maps of the reduced χ2 of different models show a linear correlation between the Mdisk and γ for the
models that have a reasonable goodness-of-fit of the radial intensity profile of 12CO. This means that we can fit the
observed 12CO intensity using either a lower disk mass with a smaller γ or higher disk mass with bigger γ. Because
the inner disk region is optically thick, and the key factor to derive a good fit is to adjust the density profile in the
outer disk region to obtain consistent intensity there.
4, The morphology of the 12CO channel maps are important constraints of the disk mass. In our parameter space,
although there are some models with higher or lower disk masses that can result in a reasonable reduced χ2, their
channel maps show a double cone structure that is either too strong or too weak compared to the ALMA observation.
The best-fit model with a disk mass of ∼ 0.1 M(cid:12) can best reproduce the observed channel maps.
This paper makes use of the following ALMA data: ADS/JAO.ALMA#2012.1.00870.S. ALMA is a partnership of
ESO (representing its member states), NSF (USA) and NINS (Japan), together with NRC (Canada), MOST and
ASIAA (Taiwan), and KASI (Republic of Korea), in cooperation with the Republic of Chile. The Joint ALMA
Observatory is operated by ESO, AUI/NRAO and NAOJ. S.J. and J.J. acknowledge support from the National
Natural Science Foundation of China (grant Nos. 11503092, 11773081, 11661161013, 11573073, 11633009), the CAS
Interdisciplinary Innovation Team, the Strategic Priority Research Program on Space Science, the Chinese Academy
of Sciences, Grant No. XDA15020302 and the Foundation of Minor Planets of Purple Mountain Observatory. A.I.
acknowledges support from the National Science Foundation through grant No. AST-1715719 and the National
Aeronautics and Space Administration support from the National Aeronautics and Space Administration through
grant No. 80HQTR18T0061 and the Center for Space and Earth Science at LANL. We thank the referee for comments
that helped to improve the manuscript.
Software: RADMC-3D v0.41(Dullemond 2012), CASA ALMA pipeline 5.4.0
ALMA Partnership, Brogan, C. L., P´erez, L. M., et al.
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., &
2015, ApJL, 808, L3
Dullemond, C. P. 2009, ApJ, 700, 1502
REFERENCES
LkCa 15 manuscript
9
Andrews, S. M., Rosenfeld, K. A., Wilner, D. J., & Bremer,
M. 2011, ApJL, 742, L5
Kraus, A. L., & Ireland, M. J. 2012, ApJ, 745, 5
Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ,
Andrews, S. M., Huang, J., P´erez, L. M., et al. 2018, ApJL,
217, 425
869, L41
Oberg, K. I., Qi, C., Fogel, J. K. J., et al. 2010, ApJ, 720,
Birnstiel, T., Dullemond, C. P., & Brauer, F. 2010, A&A,
480
513, A79
Ormel, C. W., Paszun, D., Dominik, C., et al. 2009, A&A,
Bohlin, R. C., Savage, B. D., & Drake, J. F. 1978, ApJ,
502, 845
224, 132
Oh, D., Hashimoto, J., Tamura, M., et al. 2016, PASJ, 68,
Chapillon, E., Guilloteau, S., Dutrey, A., & Pi´etu, V. 2008,
L3
A&A, 488, 565
Currie, T., Marois, C., Cieza, L., et al. 2019, ApJ, 877, L3
Dong, R., Zhu, Z., & Whitney, B. 2015, ApJ, 809, 93
Dong, R., Li, S., Chiang, E., & Li, H. 2018, ApJ, 866, 110
Draine, B. T. 2006, ApJ, 636, 1114
Dullemond, C. P., van Zadelhoff, G. J., & Natta, A. 2002,
Pi´etu, V., Dutrey, A., & Guilloteau, S. 2007, A&A, 467, 163
Pi´etu, V., Dutrey, A., Guilloteau, S., Chapillon, E., & Pety,
J. 2006, A&A, 460, L43
Pinilla, P., Benisty, M., & Birnstiel, T. 2012, A&A, 545,
A81
Pinte, C., Dent, W. R. F., Menard, F., et al. 2016, ApJ,
A&A, 389, 464
816, 25
Dullemond, C. 2012, Astrophysics Source Code Library
Dutrey, A., Wakelam, V., Boehler, Y., et al. 2011, A&A,
Pollack, J. B., Hollenbach, D., Beckwith, S., et al. 1994,
ApJ, 421, 615
535, A104
Punzi, K. M., Hily-Blant, P., Kastner, J. H., Sacco, G. G.,
Espaillat, C., Calvet, N., D'Alessio, P., et al. 2007, ApJL,
& Forveille, T. 2015, ApJ, 805, 147
670, L135
Qi, C., Kessler, J. E., Koerner, D. W., et al. 2003, ApJ,
Fedele, D., Carney, M., Hogerheijde, M. R., et al. 2017,
597, 986
A&A, 600, A72
France, K., Herczeg, G. J., McJunkin, M., et al. 2014, ApJ,
Qi, C., D'Alessio, P., Oberg, K. I., et al. 2011, ApJ, 740, 84
Rodmann, J., Henning, T., Chandler, C. J., et al. 2006,
794, 160
A&A, 446, 211
Hartmann, L., Calvet, N., Gullbring, E., & D'Alessio, P.
Rosenfeld, K. A., Andrews, S. M., Hughes, A. M., Wilner,
1998, ApJ, 495, 385
D. J., & Qi, C. 2013, ApJ, 774, 16
Henning, T., Semenov, D., Guilloteau, S., et al. 2010, ApJ,
714, 1511
Huang, J., Oberg, K. I., Qi, C., et al. 2017, ApJ, 835, 231
Huang, P., Isella, A., Li, H., Li, S., & Ji, J. 2018, ApJ, 867,
3
Huang, P., Dong, R., Li, H., Li, S., & Ji, J. 2019, in prep
Jin, S., Li, S., Isella, A., Li, H., & Ji, J. 2016, ApJ, 818, 76
Isella, A., Carpenter, J. M., & Sargent, A. I. 2009, ApJ,
701, 260
Isella, A., P´erez, L. M., & Carpenter, J. M. 2012, ApJ, 747,
Ricci, L., Testi, L., Natta, A., et al. 2010, A&A, 512, A15
Ricci, L., Testi, L., Natta, A., et al. 2010, A&A, 521, A66
Ricci, L., Mann, R. K., Testi, L., et al. 2011, A&A, 525, A81
Ricci, L., Liu, S.-F., Isella, A., & Li, H. 2018, ApJ, 853, 110
Sallum, S., Follette, K. B., Eisner, J. A., et al. 2015,
Nature, 527, 342
Simon, M., Dutrey, A., & Guilloteau, S. 2000, ApJ, 545,
1034
Skinner, S. L., & Gudel, M. 2013, ApJ, 765, 3
Thalmann, C., Grady, C. A., Goto, M., et al. 2010, ApJL,
136
718, L87
Isella, A., Chandler, C. J., Carpenter, J. M., P´erez, L. M.,
Thalmann, C., Janson, M., Garufi, A., et al. 2016, ApJL,
& Ricci, L. 2014, ApJ, 788, 129
828, L17
Isella, A., Guidi, G., Testi, L., et al. 2016, Physical Review
Thalmann, C., Mulders, G. D., Hodapp, K., et al. 2014,
Letters, 117, 251101
A&A, 566, A51
Isella, A., Huang, J., Andrews, S. M., et al. 2018, ApJL,
Thalmann, C., Mulders, G. D., Janson, M., et al. 2015,
869, L49
ApJL, 808, L41
Kenyon, S. J., & Hartmann, L. 1995, ApJS, 101, 117
Lacy, J. H., Knacke, R., Geballe, T. R., & Tokunaga, A. T.
Thi, W.-F., van Zadelhoff, G.-J., & van Dishoeck, E. F.
2004, A&A, 425, 955
1994, ApJL, 428, L69
Ubach, C., Maddison, S. T., Wright, C. M., et al. 2012,
Liu, Y., Dipierro, G., Ragusa, E., et al. 2018, A&A, 622 A75
Liu, S.-F., Jin, S., Li, S., Isella, A., & Li, H. 2018, ApJ,
857, 87
MNRAS, 425, 3137
van den Ancker, M. E., de Winter, D., & Tjin A Djie,
H. R. E. 1998, A&A, 330, 145
10
Jin et al.
van der Marel, N., van Dishoeck, E. F., Bruderer, S., P´erez,
Zhu, Z., Nelson, R. P., Hartmann, L., Espaillat, C., &
L., & Isella, A. 2015, A&A, 579, A106
Calvet, N. 2011, ApJ, 729, 47
Visser, R., van Dishoeck, E. F., & Black, J. H. 2009, A&A,
Zubko, V. G., Mennella, V., Colangeli, L., & Bussoletti, E.
503, 323
Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
1996, MNRAS, 282, 1321
LkCa 15 manuscript
11
Parameter
Mdisk (M(cid:12))
γ
RCarctan
γarctan
Table 1. Four-dimensional parameter grids
[1.0e-4, 3.3e-4, 1.0e-3, 3.3e-3, 1.0e-2, 3.3e-2, 1.0e-1, 3.3e-1]
Grids
[0.5, 1.0, 2.0, 3.0, 4.0, 5.0, 6.0, 7.0]
[1, 5, 15, 25, 35, 45, 55, 65]
[2, 4, 6, 8, 10, 12, 14, 16]
Figure 1. ALMA observations of the dust continuum, zero-moment 12CO, 13CO and C18O maps of the LkCa 15 system. Top
left: the dust continuum map. Top right: the zero-moment 12CO map. Bottom left: the zero-moment 13CO map. Bottom
right: the zero-moment C18O map.
12
Jin et al.
Figure 2. The observed channel maps of 12CO, 13CO and C18O of the LkCa 15 system.
LkCa 15 manuscript
13
Figure 3. Gas surface density profile of the best-fit run (top panel) and the surface density profile of mm-size dust used in all
the 4096 runs (bottom panel).
Figure 4. The vertical disk temperature structure of the best-fit model as calculated by our iterative approach.
10-210-1100101102Σ [ g/cm2 ]Gas10-510-410-310-210-1 0 100 200 300 400 500 600Σ [ g/cm2 ]r [ AU ]Dust14
Jin et al.
Figure 5. Comparison of the azimuthal averaged radial intensity profiles that are extracted from the zero-moment maps of
12CO, 13CO and the dust continuum map. The error-bars show the ALMA observation, and the solid green lines show our our
best-fit model. The brown and blue lines show the intensity profiles of the high-mass and low-mass model in the parameter
study at 5.2.
100101102103I (mJy/beam • km/s)ALMA 12COBest-fit modelLow-mass ModelHigh-mass Model100101102I (mJy/beam • km/s)ALMA 13COManually fit 0 2 4 6 8 10 12 14 16 0 100 200 300 400 500 600I (mJy/beam)r [ AU ]ALMA ContinuumBest-fit modelLkCa 15 manuscript
15
Figure 6. The dust continuum and zero-moment 12CO of our best-fit model and the zero-moment map of the fitted 13CO.
Left: the dust continuum map. Middle: zero-moment 12CO map. Right: zero-moment 13CO map.
Figure 7. The channel maps of 12CO of the best-fit model and the channel maps of the fitted 13CO.
16
Jin et al.
Figure 8. Heat maps of the reduced χ2 of 12CO in different combinations of parameters.
LkCa 15 manuscript
17
Figure 9. The 12CO channel maps of the low-mass, best-fit, high-mass model and the ALMA observation.
|
1604.04116 | 1 | 1604 | 2016-04-14T11:37:27 | Population synthesis of planet formation using a torque formula with dynamic effects | [
"astro-ph.EP"
] | Population synthesis studies into planet formation have suggested that distributions consistent with observations can only be reproduced if the actual Type I migration timescale is at least an order of magnitude longer than that deduced from linear theories. Although past studies considered the effect of the Type I migration of protoplanetary embryos, in most cases they used a conventional formula based on static torques in isothermal disks, and employed a reduction factor to account for uncertainty in the mechanism details. However, in addition to static torques, a migrating planet experiences dynamic torques that are proportional to the migration rate. These dynamic torques can impact on planet migration and predicted planetary populations. In this study, we derived a new torque formula for Type I migration by taking into account dynamic corrections. This formula was used to perform population synthesis simulations with and without the effect of dynamic torques. In many cases, inward migration was slowed significantly by the dynamic effects. For the static torque case, gas giant formation was effectively suppressed by Type I migration; however, when dynamic effects were considered, a substantial fraction of cores survived and grew into gas giants. | astro-ph.EP | astro-ph | Population synthesis of planet formation using a torque formula
with dynamic effects
Takanori Sasaki1 and Toshikazu Ebisuzaki2
1Department of Astronomy, Kyoto University, Kitashirakawa-Oiwake-cho, Sakyo-ku, Kyoto
606-8502, Japan
2RIKEN, 2-1 Hirosawa, Wako, Saitama 351-0198, Japan
[email protected]
Received
;
accepted
6
1
0
2
r
p
A
4
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
1
1
4
0
.
4
0
6
1
:
v
i
X
r
a
-- 2 --
ABSTRACT
Population synthesis studies into planet formation have suggested that distri-
butions consistent with observations can only be reproduced if the actual Type
I migration timescale is at least an order of magnitude longer than that deduced
from linear theories. Although past studies considered the effect of the Type
I migration of protoplanetary embryos, in most cases they used a conventional
formula based on static torques in isothermal disks, and employed a reduction
factor to account for uncertainty in the mechanism details. However, in addition
to static torques, a migrating planet experiences dynamic torques that are pro-
portional to the migration rate. These dynamic torques can impact on planet
migration and predicted planetary populations. In this study, we derived a new
torque formula for Type I migration by taking into account dynamic corrections.
This formula was used to perform population synthesis simulations with and
without the effect of dynamic torques.
In many cases, inward migration was
slowed significantly by the dynamic effects. For the static torque case, gas giant
formation was effectively suppressed by Type I migration; however, when dy-
namic effects were considered, a substantial fraction of cores survived and grew
into gas giants.
Keywords: Planetary formation -- Population synthesis -- Type I migration
-- 3 --
1.
Introduction
The population of extrasolar planets (Baruteau and Papaloizou 2013), and perhaps
even the solar system (Walsh et al. 2011), provide strong evidence that migration has played
a role in shaping planetary systems. Low-mass planets (i.e., those with masses up to that
of Neptune) migrate through the excitation of linear density waves in the disk, and through
a contribution from the corotation region (i.e., Type I migration). Early analytical work
(Tanaka et al. 2002) focused on isothermal disks, in which the temperature was prescribed
and fixed. These studies found that migration was always directed inward for reasonable
disk parameters, and that migration time scales were much shorter than the disk life time;
therefore, according to migration theory, all the planets should end up very close to the
central star.
While Type I migration has always been linked to linear interactions with the disk,
Paardekooper and Papaloizou (2009) showed that corotation torque (or horseshoe drag)
in isothermal disks show nonlinear behavior and can be much larger than previous linear
estimates, which works against fast inward migration. However,the corotation tends to be
prone to saturation and fail to prevent rapid inward migration for most of the cases, since
in the absence of a diffusive process, the corotation region is a closed system; therefore, it
can only provide a finite amount of angular momentum to a planet.
Several well-established theoretical models of planet formation based on the core
accretion scenario adopted a population synthesis approach (e.g., Ida and Lin 2004, 2008;
Ida et al. 2013; Mordasini et al. 2009a,b). Ida and Lin (2004) focused on the influence of
Type I migration on planetary formation processes and found that when the effects of Type
I migration are taken into account, planetary cores have a tendency to migrate into their
host stars before they acquire adequate mass to initiate efficient gas accretion. In order to
preserve a sufficient fraction of gas giants around solar-type stars, they introduced a Type I
-- 4 --
migration reduction factor, where factor magnitudes of smaller than unity work to lengthen
the Type I migration timescale relative to those deduced from linear theories. With a
range of small factors (∼ 0.01), it was possible to produce a planetary Mp-a distribution
that was qualitatively consistent with observations from a radial velocity survey. While
several suppression mechanisms for Type I migration under various circumstances have
been suggested (e.g., Paardekooper et al. 2011), the origin of the extremely small reduction
factor values remains unknown.
Recently, it was proposed that dynamic corotation torque can also play a role for
low-mass planets, especially where static corotation torques saturate. Paardekooper (2014)
presented an analysis of the torques on migrating, low-mass planets in locally isothermal
disks. They found that planets experience dynamic torques whenever there is a radial
gradient in vortensity in addition to static torques, which do not depend on the migration
rate. These dynamic torques are proportional to the migration rate and can have either a
positive or a negative feedback on migration, depending on whether the planet is migrating
with or against the static corotation torque. Moreover, they showed that in disks a few
times more massive than the minimum mass solar nebula (MMSN), the effects of dynamic
torques are significant to reduce inward migration.
In this study, we deduced a torque formula for Type I migration by taking into account
dynamic corrections. Using this formula, we performed population synthesis simulations
with and without the dynamic corrections in order to evaluate the migration velocity
quantitatively. We found that the effective torques with dynamic correction were much
smaller than the simple static torques when applied to disks of the MMSN model. We used
dynamic torques based on the theory of Paardekooper (2014) and estimated actual Type
I migration, and then simulated various sets of planetary systems based on the observed
range of disk properties. Finally, we compared the simulated results with observational
data.
-- 5 --
2. Dynamic Correction of Type I Migration Formula
According to Paardekooper et al. (2011; hereafter Pa11) and Colman & Nelson (2014;
hereafter CN14), new static torque formula for Type I migration can be derived (see
Appendix). When developing the dynamic correction formula, we considered the work of
Paardekooper (2014), who showed that Γdynamic, the term proportional to drp/dt, must be
included in the torque formula, or in other words:
which is given by (Pa14 eq18):
Γ = Γstatic + Γdynamic,
Γdynamic = 2π(1 − wc/w(rp))Σr2
pxsΩvp,
(1)
(2)
where Σ is the surface density of the disk, rp is the semimajor axis of the protoplanet, xs
is the thickness of the horseshoe region, Ω = (GM∗/r3
p)1/2 (G is the gravitational constant,
and M∗ is the stellar mass), and vp = drp/dt is the radial velocity of the protoplanet. Here,
1 − wc/w(rp) was calculated by (Pa14 eq28, modified by TE):
x2
s
6rpν
1 − wc/w(rp) = (3/2 + p) min(cid:18)1,
vp(cid:19) ,
(3)
where p = d ln Σ/d ln r and ν is the viscosity of the disk. Assuming a circular orbit of the
protoplanet, vp can be calculated by the equation:
τlib
dvp
dt
= −vp +
2qd
πqr3
pΩΣ
(Γstatic + Γdynamic),
(4)
where τlib = 4πrp/(3Ωxs) is the liberation timescale of gas in the disk, q = Mp/M∗, and
qd = πr2
pΣ/M∗. For the case of slow migration (i.e., τlib ≪ 4πrp/(3Ωxs) ≪ rp/vp), we were
able to assume a steady state for equation 4 to determine vp, or in other words:
− vp +
2qd
πqr3
pΩΣ
(Γstatic + Γdynamic) = 0.
(5)
-- 6 --
2.1.
Inviscid case:
x2
6rpν vp > 1
s
We derived vp by substituting equations 2 and 3 into equation 4 as:
Γinviscid =
1
1 − (3/2 + p)mc
Γstatic,
where, mc is given by (Pa14 eq20):
mc = 4qd ¯xs/q,
where ¯xs = xs/rp.
(6)
(7)
2.2. Viscid case:
x2
6rpν vp < 1
s
We derived a quadratic equation of vp by substituting equations 2 and 3 into equation
4 as:
where
A =
Av2
p + Bvp + C = 0,
2qd ¯x3
s rp
3qν (cid:18) 3
2
+ p(cid:19) τν
6rp
(8)
(9)
(10)
(11)
+ p(cid:19) = mc(cid:18)3
2
B = −1
2qdq
πh2 rpΩγstatic =
C =
rp
τmig
γstatic,
where h is the scale height of the disk, γstatic = Γstatic/Γ0 (Γ0 = (q/h)2Σr4Ω2), and τν and
τmig are the timescales of diffusion and migration, respectively, as given by (Pa14 eq10 and
23):
τν =
τmig =
r2
p ¯x2
s
ν
πh2
2qdqΩ
(12)
(13)
-- 7 --
The quadratic formula gives the total torque after dynamic correction:
Γviscid = Θ(k)Γstatic,
where the function Θ(k) is defined by (Pa14 eq30):
Θ(k) =
1 − √1 − 2k
k
,
where k is the coefficients given by (Pa14 eq31-32):
k =
8
3π(cid:18)3
2
d ¯x3
s
+ p(cid:19) γstaticq2
h2
rpΩ
ν
=(cid:18) 3
2
+ p(cid:19) mcτνγstatic
6τmig
.
(14)
(15)
(16)
The function Θ(k) takes a critical value of 2 at k = 1/2, but for k > 1/2 it does not
take any value, since the inside of the square root of Θ(k) becomes negative. This suggests
that runaway migration takes place for the case k > 1/2. Paardekooper (2014) suggested
that the time scale of migration for runaway case would be mcτmig, and in such a case,
torque for the runaway migration would be:
Γrw =
q
4qd
Γ0
(17)
The results are consistent with the numerical results of Paardekooper et al. (2011). In
their simulation, vp rapidly converged to the values obtained here, after a short transient
phase (Figs. 7, 8, and 9 in Paardekooper (2014)).
In summary, our new torque formula of Type I migration, taking into account dynamic
effects, is given as:
ΓI = Γstatic min(cid:18)
1
1 − mc(2/3 − α)
=
q
4qd
Γ0
k > 0.5
, Θ(k)(cid:19) k < 0.5
(18)
(19)
Figure 1 shows the dynamic correction factor at each semimajor axis for embryos
with masses of 0.01 M⊕, 0.1 M⊕, 1.0 M⊕, and 10 M⊕. The masses of the disks were
-- 8 --
Fig. 1. -- Dynamic correction factor at each semimajor axis for 0.01 M⊕ (M⊕ = ME), 0.1
M⊕, 1.0 M⊕, and 10 M⊕ from top to bottom in each panel. The mass of the disks are (a)
1/√10×MMSN, (b) 1.0×MMSN, (c) √10×MMSN, and (d) 10×MMSN.
-- 9 --
(a) 1/√10×MMSN, (b) 1.0×MMSN, (c) √10×MMSN, and (d) 10×MMSN. Except for a
close-in small protoplanet, most protoplanets had correction factors of significantly less
than 0.1; therefore, Type I migration was generally significantly slowed by dynamic effects.
3. Planet Formation and Migration Model
In our model, we adopted the models of Ida and Lin (2004, 2008) and Ida et al. (2013)
for (1) planetesimals' growth through cohesive collisions, (2) the evolution of planetesimal
surface density, (3) embryos' Type I migration and their stoppage at the disk inner edge
(except for a modification of the Type I migration formula to include dynamic correction;
see Section 2), and for the gas giants, (4) the onset, rate, and termination (through gap
opening and/or global depletion) of efficient gas accretion, and (5) their Type II migration.
3.1. Disk models
We adopted the MMSN model (Hayashi 1981) as a fiducial set of initial conditions for
planetesimal surface density (Σd) and introduced a multiplicative factor (fd). For the gas
surface density (Σg), we adopted the r-dependence of steady accretion disk with constant
viscosity (Σg ∝ r−1) scaled by that of the MMSN at 10 AU with a scaling factor (fg).
Following Ida and Lin (2008), we set:
Σd = Σd,10ηicefd(r/10AU)−1.5,
Σg = Σg,10fg(r/10AU)−1.0,
(20)
where normalization factors Σd,10 = 0.32 g cm−2 and Σg,10 = 75 g cm−2, and the step
function was ηice = 1 inside the ice line at aice and 4.2 for r > aice.
Neglecting the detailed energy balance in the disk (Chiang and Goldreich 1997), we
adopted the equilibrium temperature distribution of optically thin disks given by Hayashi
-- 10 --
(1981), such that:
T = 280(cid:16) r
1AU(cid:17)−1/2(cid:18) L∗
L⊙(cid:19)1/4
K,
(21)
where L∗ and L⊙ are stellar and solar luminosity. We set the ice line to be that determined
by an equilibrium temperature in optically thin disk regions:
aice = 2.7(cid:18) L∗
L⊙(cid:19)1/2
AU.
Owing to viscous diffusion and photoevaporation, fg decreases with time. For
simplicity, we adopted:
where fg,0 is the initial value of fg and τdep is the gas depletion timescale.
fg = fg,0 exp(cid:18)−
t
τdep(cid:19) ,
(22)
(23)
3.2. From oligarchic growth to isolation
On the basis of the oligarchic growth model (Kokubo and Ida 1998, 2002), the growth
rate of embryos/cores at any location, a, and time t, in the presence of disk gas, was
described by:
where
dMc
dt
=
Mc
τc,acc
(24)
yr,
(25)
τc,acc = 3.5 × 105η−1
ice f −1
d f −2/5
g
1AU(cid:17)5/2(cid:18) Mc
(cid:16) a
M⊕(cid:19)1/3(cid:18) M∗
M⊙(cid:19)−1/6
where Mc is the mass of the embryo/core. Furthermore, we set the mass of typical field
planetesimals to be m = 1020 g.
We computed the evolution of Σd distribution due to accretion by all emerging embryos
in a self-consistent manner. The growth and migration of many planets were integrated
simultaneously with the evolution of the Σd-distribution.
-- 11 --
During the early phase of evolution, embryos are embedded in their natal disks.
Despite their mutual gravitational perturbation, embryos preserve their circular orbits
owing to gravitational drag from disk gas (Ward 1993) and dynamic friction from residual
planetesimals (Stewart and Ida 2000). After the disk gas is severely depleted, the efficiency
of the eccentricity damping mechanism is reduced, and the embryos' eccentricity grow until
they cross each other's orbits (i.e., giant impact). However, in this study, growth via the
giant impact process was not considered. Moreover, we also ignored dynamic interaction
between planets, with the growth of individual planets integrated independently.
3.3. Type I migration
Type I migration of an embryo is caused by the sum of tidal torque from disk regions
that are both interior and exterior to the embryos. The rate and direction of embryos'
migration are determined by the differential Lindblad and corotation torques. While a
conventional formula of Type I migration, which assumes locally isothermal disks (Tanaka
et al. 2002), shows that the migration is always inward, recent developments have shown
Type I migration of isolated embryos in non-isothermal disks; therefore, the magnitude and
sign of tidal torque can be changed. Ida and Lin (2008) used the conventional formula of
Type I migration in isothermal disks derived by Tanaka et al. (2002) with a scaling factor
C1 of:
dr
dt ≃ C1 × 1.08(p + 0.80q − 2.52)
Mp
M∗
Σgr2
cs (cid:19)2
M∗ (cid:18)rΩK
rΩK,
(26)
where p = d log Σg/d log r, q = d log T /d log r, cs is the sound speed, and ΩK is the Keplerian
angular velocity. The expression of Tanaka et al. (2002) corresponds to C1 = 1, and for
slower migration, C1 < 1. While we derived a new torque formula for Type I migration that
included dynamic corrections (see Section 2), for comparison we also used the conventional
formula with the scaling factor C1 = 1.0.
-- 12 --
We assumed that Type I migration ceases inside the inner boundary of the disk,
because at this point fg is locally zero. For computational convenience, we set the disk
inner boundary to be the edge of the magnetospheric cavity at 0.04 AU.
3.4. Formation of gas giant planets
Models for the formation of gas giant planets were the same as those used in Ida et al.
(2013). Embryos were surrounded by gaseous envelopes when their surface escape velocities
became larger than the sound speed of the surrounding disk gas. When their mass grew
(through planetesimal bombardment) above a critical mass:
Mc,hydro ≃ 10
Mc
10−6M⊕yr−1!0.25
M⊕,
(27)
both the radiative and convective transport of heat became sufficiently efficient to allow
their envelope to contract dynamically (Ikoma et al. 2000).
In the above equation, we neglected the dependence on opacity in the envelope (Hori
and Ikoma 2010). In regions where cores have already acquired isolation mass, their
planetesimal-accretion rate ( Mc) would be much diminished (Ikoma et al. 2000) and Mc,hydro
would be comparable to an Earth-mass, M⊕. However, gas accretion also releases energy
and its rate is still regulated by the efficiency of radiative transfer in the envelope, such
that:
dMp
dt ≃
Mp
τKH
,
(28)
where Mp is the planet mass including gas envelope. According to Ida and Lin (2008), we
approximated the Kelvin-Helmholtz contraction timescale, τKH, of the envelope using:
τKH ≃ 109(cid:18) Mp
M⊕(cid:19)−3
,
(29)
Equation 28 shows that dMp/dt rapidly increases as Mp grows; however, this is limited
-- 13 --
by the global gas accretion rate throughout the disk and by the process of gap formation
near the protoplanets' orbits. The disk accretion rate can be expressed as:
Mdisk ≃ 3 × 10−9fg(cid:16) α
10−3(cid:17) M⊙yr−1,
(30)
where α is a parameter of alpha prescription for turbulent viscosity (Shakura and Sunyaev
1973). During the advanced stage of disk evolution, we assumed that both Mdisk and Σg
evolved in proportion to exp(−t/τdep). The rate of accretion onto the cores cannot exceed
Mdisk.
A gap, or at least a partial gap, is formed when a planet's tidal torque exceeds the
disk's intrinsic viscous stress (Lin and Papaloizou 1985). This viscous condition for gap
formation is satisfied for planets with:
Mp > Mg,vis ≃ 30(cid:16) α
10−3(cid:17)(cid:16) a
1AU(cid:17)1/2(cid:18) L∗
L⊙(cid:19)1/4
M⊕.
(31)
In this case, Type I migration transitions to Type II migration. Fluid dynamic simulations
(D'Angelo and Kley 2003; D'Angelo and Lubow 2008) show that some fraction of gas
still flows into the gap. Following the results of Dobbs-Dixon et al. (2007), we completely
terminated gas accretion when a planet's Hill radius became larger than two times the disk
scale height, which corresponded to the thermal condition of (Lin and Papaloizou 1985),
that is:
Mp > Mg,th ≃ 0.95 × 103(cid:16) a
1AU(cid:17)3/4(cid:18) L∗
L⊙(cid:19)3/8(cid:18) M∗
M⊙(cid:19)−1/2
M⊕.
(32)
In general, our models for gas accretion rates onto the cores were:
dMp
dt
= fgap Mp,nogap,
(33)
when in the absence of any feedback on the disk structure. Therefore, without the effect of
gap opening:
Mp,nogap = min(cid:18) Mp
τKH
,
Mdisk(cid:19) ,
(34)
-- 14 --
and fgap is a reduction factor due to gap opening:
1
[for Mp < Mg,vis]
log Mp − log Mg,vis
log Mg,th − log Mg,vis
0
[for Mg,vis < Mp < Mg,th],
(35)
[for Mp > Mg,th].
fgap =
3.5. Type II migration
During gap formation, embedded gas giants adjust their positions in the gap to
establish a quasi equilibrium between the torque applied on them from disk regions both
interior and exterior to their orbits. Subsequently, as the disk gas undergoes viscous
diffusion, this interaction leads to Type II migration.
We assumed that planets undergo Type II migration after they have accreted a
sufficient mass to satisfy the viscous condition (Mg,vis < Mp) for gap formation.
While Mp increases, the disk mass declines owing to stellar and planetary accretion
and photoevaporation. While disk mass exceeds Mp (during the disk-dominated regime),
planets' Type II migration is locked, with the viscous diffusion of the disk gas. During
the advanced stages of disk evolution, when the mass becomes smaller than Mp (during
the planet-dominated regime), embedded planets carry a major share of the total angular
momentum content.
For the disk-dominated regime, the migration timescale is given by:
τmig2,disk ≃ 0.7 × 105(cid:16) α
10−3(cid:17)−1(cid:16) a
1AU(cid:17)(cid:18) M∗
M⊙(cid:19)−1/2
yr.
(36)
For the planet-dominated regime, the migration timescale is given by:
τmig2,pl ≃ 5 × 105f −1
g (cid:18) C2α
10−4(cid:19)−1(cid:18) Mp
MJ(cid:19)(cid:16) a
1AU(cid:17)1/2(cid:18) M∗
M⊙(cid:19)−1/2
yr,
(37)
-- 15 --
where C2 is an efficiency factor associated with the degree of asymmetry in the torques
between the inner and outer disk regions. If the inner disk is severely depleted, C2 = 1. We
treated the factor C2 as a model parameter, and we set C2 = 0.1.
4. Population Synthesis of Planetary Systems
Using our new torque formula for Type I migration, we modeled the formation
of planetary systems using Monte Carlo simulations. The predicted mass and period
distributions were compared with those from a conventional Type I migration model.
4.1. Numerical settings
We first generated a set of 1000 disks with various values of fg,0 (the initial value of
fg) and τdep. We adopted a range of disk model parameters that represented the observed
distribution of disk properties and assigned them to each model with an appropriate
statistical weight. For the gaseous component, we assumed that fg,0 had a lognormal
distribution centered on fg,0 = 1 with a dispersion of 1 and an upper cutoff at fg,0 = 30,
independent of the stellar metallicity. For heavy elements, we choose fd,0 = 10[Fe/H]dfg,0,
where [Fe/H]d is the metallicity of the disk. We assumed that these disks had the same
metallicities as their host stars. We also assumed that τdep had log-uniform distributions in
the range 106 − −107 yr.
For each disk, 15 values of a for the protoplanetary seeds were selected from a
long-uniform distribution in the range 0.05 -- 30 AU, assuming that the mean orbital
separation between planets was 0.2 on a logarithmic scale. Constant spacing in the
logarithm corresponded to the spacings between the cores, which were proportional to a.
This represented the simplest choice and a natural outcome of dynamic isolation at the end
-- 16 --
of oligarchic growth.
In all simulations, the values α = 10−3 and M∗ = 1M⊙ were assumed. Since ongoing
radial velocity surveys are focused on relatively metal-rich stars, we presented our results
with [Fe/H] = 0.1.
We artificially terminated Type I and Type II migration near the disk inner edge at
0.04 AU. We did not specify a survival criterion for the close-in planets because we lacked
adequate knowledge about planets' migration and their interaction with host stars near
the inner edge of their nascent disks. Hence, we recorded all of the planets that migrated
to the vicinity of their host stars. In reality, a large fraction of the giant planets that
have migrated to small disk radii were either consumed (Sandquist et al. 1998) or tidally
disrupted (Trilling et al. 1998) by their host stars. Cores that migrate to the inner edge of
the disk may also coagulated and form super-Earths (Ogihara and Ida 2009); however, this
was not considered in our simulations.
4.2. Simulated individual systems
We compared the time evolution of planetary masses and semimajor axes for the new
torque formula model (Fig. 2a; Fig. 3a) and the conventional torque formula model with
C1 = 1.0 (Fig. 2b; Fig. 3b). We choose a disk a few times more massive (fg,0 = 6.0 and
8.0) than the minimum solar nebula, and with τdep = 3 × 106 yr. The results showed that
inward migration of planet embryos was slowed significantly by dynamic effects. When
dynamic effects were considered, some cores survived and grew into gas giants; however,
when considering only static torque, all cores migrated to the vicinity of their central star
before growing enough to accrete the nebula gas.
-- 17 --
Fig. 2. -- Growth and migration of planets for scaling factor fg,0 = 3.0. Units of mass (Mp)
and semimajor axis (a) are Earth masses (M⊕ = ME) and AU. (a) Mass evolutions obtained
from simulations with the new torque formula for Type I migration. (b) Evolutions using
the conventional formula (C1 = 1.0).
Fig. 3. -- Growth and migration of planets for scaling factor fg,0 = 8.0. Units of mass (Mp)
and semimajor axis (a) are Earth masses (M⊕ = ME) and AU. (a) Mass evolutions obtained
from simulations with the new torque formula for Type I migration. (b) Evolutions using
the conventional formula (C1 = 1.0).
-- 18 --
Fig. 4. -- Evolution of the dynamic correction factor for a scaling factor of (a) fg,0 = 3.0,
and (b) fg,0 = 8.0. Units of mass (Mp) are Earth masses (M⊕ = ME).
Figure 4 shows the evolution of the dynamic correction factor (see equation 18),
ΓI/Γstatic = min(cid:18)
1
1 − mc(3/2 + p)
, Θ(k)(cid:19) ,
(38)
with the mass of each planet embryo. The correction factors remained small (≤ 0.1)
throughout the simulation; therefore, the dynamic correction of Type I migration effectively
prevented embryos from migrating to the central star.
4.3. Distributions of mass and semimajor axes
We compared the predicted Mp-a distributions using the new torque formula (Fig. 5a)
and the conventional formula with C1 = 1.0 at t = 2 × 107 yr. In order to directly compare
the theoretical predictions with the observed data, we plotted values of Mp that were 1.27
times the values of Mp sin i, as determined from radial velocity measurements (Fig. 5c).
This correction factor corresponded to mean values of 1/hsin ii = 4/π for a sample of
planetary systems with randomly oriented orbital plants. To compare the theoretical results
-- 19 --
with M∗ = 1M⊙, we plotted only the data of planets around stars with M∗ = 0.8 − 1.2M⊙
that have been observed by radial velocity surveys 1.
For the conventional models, the formation probability of gas giants dramatically
changed with C1 (Ida and Lin 2008). Within the limits of Type I migration with an
efficiency comparable to that deduced from the traditional linear torque analysis (i.e.,
with C1 = 1; Fig. 5b), all cores were cleared prior to gas depletion, such that gas giant
formation was effectively suppressed. However, when considering the dynamic effects, a
substantial fraction of the cores survived and grew into large gas giants (Fig. 5a). We
carry out a Kolmogorov-Smirnov (K-S) test for statistical similarity between the predicted
Mp-a distributions and the observed data for the parameter domain of 0.1 AU < a < 5 AU
and Mp > 100M. While the conventional model produces no giant planets (Fig. 5b), the
predicted Mp-a distribution using the new torque formula (Fig. 5a) is statistically similar
to the observed data (Fig. 5c) within a significance level of p-value > 0.05 for both the
semimajor axis and mass cumulative distribution functions.
Without considering the dynamic correction for Type I migration, when a planet's
mass exceeded that of Earth, the corotation torque became smaller owing to saturation. At
this point, static torque affected the planet more efficiently so that its inward migration was
rapid. However, when the dynamic correction was included, the migration timescale was
short, and the saturation of the corotation torque was less effective. Under these conditions,
inward migration slowed, which allowed for the formation of gas giants before migration to
the central star.
In summary, population synthesis simulations using our new torque formula with
dynamic correction (Fig. 5a) can explain the gas giants (> 100M⊕) observed in exoplanetary
1See http://exoplanet.eu/.
-- 20 --
systems (Fig. 5c). In contrast, simulations using a conventional formula (C1 = 1.0; Fig. 5b)
cannot explain the observed data. These results show that planet populations consistent
with observations can be reproduced naturally (i.e., without considering the reduction
factor) if we take into account dynamic corrections for Type I migration torque.
5. Conclusions
We derived a new torque formula for Type I migration by taking into account dynamic
corrections. Using this formula, we performed population synthesis simulations with and
without the effects of dynamic torques. In most cases, inward migration was significantly
slowed by the dynamic effects. Considering just static torques, gas giant formation was
effectively suppressed by Type I migration of cores; however, when dynamic effects were
considered, a substantial fraction of cores survived and grew into gas giants.
Acknowledgments
We thank an anonymous reviewer for the helpful comments. This research was
supported by Grant-in-Aid for Scientific Research on Innovative Areas from the Ministry of
Education, Culture, Sports, Science and Technology (MEXT; Grant Number 26106006).
T.S. was supported by a Grand-in-Aid for Young Scientists (KAKENHI B) from the Japan
Society for the Promotion of Science (JSPS; Grant Number 24740120).
-- 21 --
Fig. 5. -- Planetary mass and semimajor axis distribution. Units of mass (Mp) and semimajor
axis (a) are Earth masses (M⊕ = ME) and AU. (a) Distribution obtained from Monte Carlo
simulations with the new torque formula for Type I migration. (b) Distributions using the
conventional formula (C1 = 1.0). (c) Observational data of extrasolar planets around stars
with M∗ = 0.8−1.2M⊙ detected by radial velocity surveys. The determined value of Mp sin i
is multiplied by 1/hsin ii = 4/π ≃ 1.27, where a random orientation of the planetary orbital
planes is assumed.
-- 22 --
REFERENCES
Baruteau, C., Papaloizou, J.C.B., 2013. Disk-planets interactions and the diversity of period
ratios in Kepler's multi-planetary systems. Astrophysical Journal 778, 7 (15 pp).
Chiang, E.I., Goldreich, P., 1997. Spectral energy distributions of T Tauri stars with passive
circumstellar disks. Astrophysical Journal 490, 368 -- 376.
Coleman, G.A.L., Nelson, R.P., 2015. On the formation of planetary systems via
oligarchic growth in thermally evolving viscous discs. Monthly Notices of the Royal
Astronomical Society 445, 479 -- 499.
Cresswell, P., Nelson, R.P., 2008. Three-dimensional simulations of multiple protoplanets
embedded in a protostellar disc. Astronomy & Astrophysics 482, 677 -- 690.
D'Angelo, G., Kley, W., 2003. Orbital migration and mass accretion of protoplanets in
three-dimensional global computations with nested grids. Astrophysical Journal 586,
540 -- 561.
D'Angelo, G., Lubow, S.H., 2008. Evolution of migrating planets undergoing gas accretion.
Astrophysical Journal 685, 560 -- 583.
Dobbs-Dixon, I., Li, S.L., Lin, D.N.C., 2007. Tidal barrier and the asymptotic mass of
proto gas-giant planets. Astrophysical Journal 660, 791 -- 806.
Hayaschi, C., 1981. Structure of the solar nebula. Growth and decay of magnetic fields, and
effect of magnetic and turbulent viscosities on the nebula. Progress of Theoretical
Physics Supplement 70, 35 -- 53.
Hori, Y., Ikoma, M., 2010. Critical core masses for gas giant formation with grain-free
envelopes. Astrophysical Journal 714, 1343 -- 1346.
-- 23 --
Ida, S., Lin, D.N.C., 2004. Toward a deterministic model of planetary formation. I. A desert
in the mass and semimajor axis distributions of extrasolar planets. Astrophysical
Journal 604, 388 -- 413.
Ida, S., Lin, D.N.C., 2008. Toward a deterministic model of planetary formation. IV. Effects
of type I migration. Astrophysical Journal 673, 487 -- 501.
Ida, S., Lin, D.N.C., Nagasawa, M., 2013. Toward a deterministic model of planetary
formation. VII. Eccentricity distribution of gas giants. Astrophysical Journal 775, 42
(25 pp).
Ikoma, M., Nakazawa, K., Emori, H., 2000. Formation of giant planets: Dependences on
core accretion rate and grain opacity. Astrophysical Journal 537, 1013 -- 1025.
Kokubo, E., Ida, S., 1998. Oligarchic growth of protoplanets. Icarus 131, 171 -- 178.
Kokubo, E., Ida, S., 2002. Formation of protoplanet systems and diversity of planetary
systems. Astrophysical Journal 581, 666 -- 680.
Lin, D.N.C., Papaloizou, J.C.B., 1985. On the dynamical origin of the solar system. In:
Black, D.C., Matthew, M.S. (Eds.), Protostars and Planets II. Univ. Arizona Press,
Tucson, pp. 981 -- 1072.
Mordasini, C., Alibert, Y., Benz, W., 2009. Extrasolar planet population syntehsis. I.
Method, formation tracks, and mass-distance distribution. Astronomy & Astrophysics
501, 1139 -- 1160.
Mordasini, C., Alibert, Y., Benz, W., Naef, D., 2009. Extrasolar planet population
syntehsis. II. Statistical comparison with observations. Astronomy & Astrophysics
501, 1161 -- 1184.
-- 24 --
Ogihara, M., Ida, S., 2009. N-body simulations of planetary accretion around M dwarf
stars. Astrophysical Journal 699, 824 -- 838.
Paardekooper, S.-J., Papaloizou, J.C.B., 2009. On corotation torques, horseshoe drag and
the possibility of sustained stalled or outward protoplanetary migration. Monthly
Notices of the Royal Astronomical Society 394, 2283 -- 2296.
Paardekooper, S.-J., Baruteau, C., Kley, W., 2011. A torque formula for non-isothermal
type I planetary migration - II. Unsaturated horseshoe drag. Monthly Notices of the
Royal Astronomical Society 410, 293 -- 303.
Paardeokooper, S.-J., 2014. Dynamical corotation torque on low-mass planets. Monthly
Notices of the Royal Astronomical Society 444, 2031 -- 2042.
Sandquist, E., Taam, R.E., Lin, D.N.C., Burkert, A., 1998. Planet consumption and stellar
metallicity enhancements. Astrophysical Journal 506, L65 -- L68.
Shakura, N.I., Sunyaev, R.A., 1973. Black holes in binary systems. Observational
appearance. Astronomy & Astrophysics 24, 337 -- 355.
Stewart, G.R., Ida, S., 2000. Velocity evolution of planetesimals: Unifield analytical
formulas and comparisons with N-body simulations. Icarus 143, 28 -- 44.
Tanaka, H., Takeuchi, T., Ward, W.R., 2002. Three-dimensional interaction between a
planet and an isothermal gaseous disk. I. Corotation and Lindblad torques and
planet migration. Astrophysical Journal 565, 1257 -- 1274.
Trilling, D.E., Benz, W., Guillot, T., Lunine, J.I., Hubbard, W.B., Burrows, A., 1998.
Orbital evolution and migration of giant planets: Modeling extrasolar planets.
Astrophysical Journal 500, 428 -- 439.
-- 25 --
Walsh, K.J., Morbidelli, A., Raymond, S.N., O'Brien, D.P., Mandel, A.M., 2011. A low
mass for Mars from Jupiter?s early gas-driven migration. Nature 475, 206 -- 209.
Ward, W.R., 1993. Density waves in the solar nebula: Planetesimal velocities. Icarus 106,
274 -- 287.
A. Static Torque Formula
Static torque (Pa11 and CN14) is given by
Γstatic = FLΓLR
+ (cid:20)ΓVHSF (pν)G(pν) + ΓEHSF (pν)F (pχ)qG(pν)G(pχ)
+ ΓLVCT(1 − K(pν)) +ΓLECTq(1 − K(pν))(1 − K(pχ))(cid:21) FeFi,
(A1)
(A2)
(A3)
where ΓLR, ΓVHS, ΓEHS, ΓLVCT, and ΓLECT are the Lindblad torque, vortensity and entropy
related horseshoe drag torques, and linear vortensity and entropy related corotation torques,
respectively, as given by equation 3 -- 7 in Paardekooper et al. (2011):
ΓLR = (−2.5 − 1.7β + 0.1α)Γ0/γeff,
ΓVHS = [1.1(3/2 − α)]Γ0/γeff,
ΓEHS = 7.9(ξ/γeff)Γ0/γeff,
ΓLVCT = [0.7(3/2 − α)]Γ0/γeff,
ΓLECT = [(2.2 − 1.4/γeff)]Γ0/γeff,
(A4)
(A5)
(A6)
(A7)
(A8)
where α = d ln Σ/d ln r, β = d ln Tm/d ln r, and ξ = β − (γeff − 1)α. Here, Γ0 = (q/h)2Σr4Ω2.
This manuscript was prepared with the AAS LATEX macros v5.2.
-- 26 --
The functions F (pν), F (pχ), G(pν), F (pχ), K(pν), and K(pχ) are related to the ratio
between the viscous/thermal diffusion time scale and horseshoe liberation/horseshoe U-turn
time scales, given by equations 23, 30, and 21 in Paardekooper et al. (2011):
F (p) =
1
1 + (p/1.3)2
G(p) =
16
8 (cid:1)3/4 p3/2
25(cid:0) 45π
45π(cid:1)4/3
1 − 9
25(cid:0) 8
p−8/3
45π
p <q 8
p >q 8
45π
K(p) =
16
28 (cid:1)3/4 p3/2
25(cid:0) 45π
45π(cid:1)4/3 p−8/3
25(cid:0) 28
1 − 9
45π
p <q 28
p >q 28
45π .
The pν and pχ are given by (Pa11 eq19 and eq40):
s
2
pν =
3r r2Ω¯x3
pχ = sr2Ω¯x3
2πχ
2πν
,
s
where (Pa11 eq48-49) ¯xs = xs/r is given by
and (Pa11 eq34)
1.1
γ1/4
¯xs = Cpq/h,
b/h(cid:19)−1/4
eff (cid:18) 0.4
4γ(γ − 1)σT 4
3κρ2H 2Ω2
.
,
C =
χ =
(A9)
(A10)
(A11)
(A12)
(A13)
(A14)
(A15)
(A16)
(A17)
(A18)
-- 27 --
The effective adiabatic index γeff is given by (Pa11 eq45-46):
γeff =
2Qγ
γQ + 1
2q2p(γ2Q2 + 1)2 − 16Q2(γ − 1) + 2γ2Q2 − 2
taking into account of the photon diffusion in a disk. Here,
,
(A19)
Q =
2χ
3h3r2Ω
(A20)
Fe and Fi are the reduction factor due to eccentricity and inclination of the planets,
which are give by (CN15 eq16-20):
Fe = exp(cid:18)−
e
ef(cid:19) ,
where e is the plane's eccentricity and ef is defined as:
ef = h/2 + 0.01
Fi = 1 − tanh(i/h),
(A21)
(A22)
(A23)
where i is the inclination of the planet. The factor FL is the reduction in Lindblad torques
when planets are on eccentric or inclined orbits, and is given by Cresswell and Nelson
(2008):
FL ="Pe +
where Pe is defined as
Pe =
Pe
Pe × 0.07(cid:18) i
e(cid:19)4
h(cid:19) + 0.085(cid:18) i
h(cid:19)2!#−1
h(cid:17)(cid:18) i
− 0.08(cid:16) e
2.25h(cid:1)1/2 +(cid:0) e
2.84h(cid:1)6
2.02h(cid:1)4
1 −(cid:0) e
1 +(cid:0) e
,
(A24)
(A25)
|
1607.06686 | 2 | 1607 | 2016-08-05T18:10:29 | The analemma criterion: accidental quasi-satellites are indeed true quasi-satellites | [
"astro-ph.EP"
] | In the Solar system, a quasi-satellite is an object that follows a heliocentric path with an orbital period that matches almost exactly with that of a host body (planetary or not). The trajectory is of such nature that, without being gravitationally attached, the value of the angular separation between host and quasi-satellite as seen from the Sun remains confined within relatively narrow limits for time-spans that exceed the length of the host's sidereal orbital period. Here, we show that under these conditions, a quasi-satellite traces an analemma in the sky as observed from the host in a manner similar to that found for geosynchronous orbits. The analemmatic curve (figure-eight-, teardrop-, ellipse-shaped) results from the interplay between the tilt of the rotational axis of the host and the properties of the orbit of the quasi-satellite. The analemma criterion can be applied to identify true quasi-satellite dynamical behaviour using observational or synthetic astrometry and it is tested for several well-documented quasi-satellites. For the particular case of 15810 (1994 JR1), a putative accidental quasi-satellite of dwarf planet Pluto, we show explicitly that this object describes a complex analemmatic curve for several Plutonian sidereal periods, confirming its transient quasi-satellite status. | astro-ph.EP | astro-ph |
MNRAS 000, 1 -- 6 (2016)
Preprint 13 October 2018
Compiled using MNRAS LATEX style file v3.0
The analemma criterion: accidental quasi-satellites are indeed true
quasi-satellites
C. de la Fuente Marcos⋆ and R. de la Fuente Marcos
Apartado de Correos 3413, E-28080 Madrid, Spain
Accepted 2016 July 22. Received 2016 July 22; in original form 2016 July 1
ABSTRACT
In the Solar system, a quasi-satellite is an object that follows a heliocentric path with an orbital
period that matches almost exactly with that of a host body (planetary or not). The trajectory is
of such nature that, without being gravitationally attached, the value of the angular separation
between host and quasi-satellite as seen from the Sun remains confined within relatively nar-
row limits for time-spans that exceed the length of the host's sidereal orbital period. Here, we
show that under these conditions, a quasi-satellite traces an analemma in the sky as observed
from the host in a manner similar to that found for geosynchronous orbits. The analemmatic
curve (figure-eight-, teardrop-, ellipse-shaped) results from the interplay between the tilt of the
rotational axis of the host and the properties of the orbit of the quasi-satellite. The analemma
criterion can be applied to identify true quasi-satellite dynamical behaviour using observa-
tional or synthetic astrometry and it is tested for several well-documented quasi-satellites. For
the particular case of 15810 (1994 JR1), a putative accidental quasi-satellite of dwarf planet
Pluto, we show explicitly that this object describes a complex analemmatic curve for several
Plutonian sidereal periods, confirming its transient quasi-satellite status.
Key words: methods: numerical -- celestial mechanics -- minor planets, asteroids: general --
minor planets, asteroids: individual: 15810 (1994 JR1) -- minor planets, asteroids: individual:
63252 (2001 BL41).
1 INTRODUCTION
Objects trapped in a 1:1 mean motion resonance with a host (plan-
etary or not) are classified as co-orbitals of the host, independently
of the shape and orientation of their paths (Morais & Morbidelli
2002); in other words, to be classed as co-orbitals their orbits do
not have to resemble that of the host as long as the ratio of their
orbital periods equates to almost exactly one. In general, co-orbital
configurations are not identified observationally but as a result of
the statistical analysis of large sets of numerical integrations. There
is, however, a potential exception to this standard approach; a par-
ticular type of co-orbital configuration that can be confirmed obser-
vationally, the quasi-satellite dynamical state.
Here, we study the apparent motion in host-centric equato-
rial coordinates of known quasi-satellites to show that they trace an
analemmatic curve in the sky as observed from the host in a manner
similar to that found for geosynchronous orbits. This paper is orga-
nized as follows. Section 2 discusses the so-called analemma crite-
rion for quasi-satellites and it includes an extensive exploration of
the known quasi-satellite population. The particular case of 15810
(1994 JR1), a putative accidental quasi-satellite of Pluto, is anal-
ysed in Section 3 to show that according to the analemma criterion
⋆ E-mail: [email protected]
c(cid:13) 2016 The Authors
it is a true transient quasi-satellite of Pluto. Results are discussed in
Section 4 and conclusions are summarized in Section 5.
2 THE ANALEMMA CRITERION
Minor bodies are confirmed as co-orbitals after statistical analysis
of their simulated orbital evolution. Here, we show that there is a
particular type of co-orbital configuration that may be confirmed
observationally: the quasi-satellite state.
2.1 Quasi-satellites: a short review and a lost specimen
Minor bodies engaged in quasi-satellite behaviour with a host move
near the host for the duration of the quasi-satellite episode although
each pair minor-body -- host is not gravitationally bound. In the Solar
system and from a frame of reference centred at the Sun but coro-
tating with the host, the quasi-satellite appears to go around the host
like a regular retrograde satellite but the physical distance between
the two bodies is always greater than the radius of the Hill sphere of
the host (see e.g. fig. 1 in Mikkola et al. 2004). The quasi-satellite
state is one of the dynamical epitomes of the 1:1 mean motion or
co-orbital resonance, the other two being the Trojan or tadpole and
the horseshoe resonant states (see e.g. Murray & Dermott 1999;
Mikkola et al. 2006).
2
C. de la Fuente Marcos and R. de la Fuente Marcos
Dynamical classification within the 1:1 mean motion reso-
nance is based on the study of a critical angle, the relative mean
longitude, λr, or difference between the mean longitude of the ob-
ject and that of its host. If λr librates or oscillates over time, then the
object under study is a co-orbital. In principle, this can only be con-
firmed via N-body simulations. Quasi-satellites exhibit libration of
λr about 0◦(for additional details, see e.g. Mikkola et al. 2006; de
la Fuente Marcos & de la Fuente Marcos 2014, 2016a,b).
The existence of quasi-satellites was predicted more than a
century ago (Jackson 1913), but the first bona fide quasi-satellite
was not identified until much later -- 2002 VE68 was confirmed as
quasi-satellite of Venus by Mikkola et al. (2004). However, the first
quasi-satellite may have been identified in 1973 although it was
apparently lost shortly after. Using numerical integrations, Cheb-
otarev (1974)1 showed that the so-called minor planet 7617 (see
his fig. 5 and table 5) was a quasi-satellite of Jupiter although
he regarded this object as a distant Jovian satellite. This minor
planet 7617 is clearly (see table 5 in Chebotarev 1974) not aster-
oid 7617 (1996 VF30), as Chebotarev (1974) followed the number-
ing scheme in van Houten et al. (1970). The orbital elements (1950
equinox) of the mysterious minor planet 7617 in van Houten et al.
(1970) -- a = 5.0785 au, e = 0.6179, i = 4.◦080, Ω = 68.◦81 and
ω = 209.◦23 -- do not match those of any known asteroid or comet;
therefore, it is pressumed lost.
2.2 Theoretical expectations
When observed from a celestial object (planetary or not) true satel-
lites (not following synchronous orbits), co-orbitals of the Trojan
or horseshoe type, and passing objects describe roughly sinusoidal
paths in the sky over a sidereal orbital period. In sharp contrast,
quasi-satellites appear to orbit the host when viewed in a heliocen-
tric frame of reference that rotates with the host. As their orbital
periods are very close to the sidereal period of the host, the stan-
dard sinusoidal trace becomes compressed longitudinally turning
into an analemmatic curve.
The analemma or analemmatic curve -- the figure-eight
loop -- has been traditionally linked to graphic depictions of the
changing of the seasons and the equation of time (see e.g. Heath
1923; Raisz 1942; Oliver 1972; di Cicco 1979; Irvine 2001; Hol-
brow 2013). In addition, the trajectories of geosynchronous satel-
lites as observed from the ground have the appearance of an
analemma (Chalmers 1987). From the host, the apparent motion of
a quasi-satellite during a sidereal orbital period is not too different
from that of true satellites moving in synchronous orbits.
In order to show that a given quasi-satellite traces an analem-
matic curve, we proceed as follows. We perform full N-body sim-
ulations in ecliptic coordinates; at time t, for a given host of co-
ordinates, (xh, yh, zh), and axial tilt or obliquity, ǫ, and a certain
quasi-satellite located at (xqs, yqs, zqs) we can define the host-centric
equatorial coordinates, (α∗, δ∗):
r cos α∗ cos δ∗ = xqs − xh
r sin α∗ cos δ∗ = (yqs − yh) cos ǫ − (zqs − zh) sin ǫ
r sin δ∗ = (yqs − yh) sin ǫ + (zqs − zh) cos ǫ ,
(1)
where r is the distance between host and quasi-satellite at time t.
For the particular case of the Earth, (α∗, δ∗) become (α, δ), the usual
geocentric equatorial coordinates. Over one sidereal period, there is
a north-south oscillation of δ∗ that is responsible for the lengthwise
extension of the analemma pattern. Such libration is induced by
the fact that the orbital plane of the quasi-satellite and the celestial
equator at the host are, in general, tilted by a certain amount. In
addition, the relative motion of a quasi-satellite with respect to its
host is not uniform because, in a typical case, their orbital eccen-
tricities are different although their semimajor axes are nearly equal
and this tends to distort the analemma. The observed apparent mo-
tion results from the interplay between the two effects; when both
have comparable strengths, the familiar figure-eight is obtained. In
the particular case of the Earth and for an ideal quasi-satellite bright
enough to be observed year-round with standard ground-based tele-
scopes, regular astrometric observations should make it possible
to plot the associated analemmatic curve without any help from
numerical computations. Unfortunately, no such quasi-satellite is
known to exist and N-body simulations are needed to produce syn-
thetic astrometry to confirm the theoretical expectations.
The apparent motion of the objects studied here and plotted in
Figs 1 and 2 has been computed using the Hermite scheme (Makino
1991; Aarseth 2003).2 The Cartesian state vectors of the integrated
bodies at the epoch 2457600.5 JD TDB (2016-July-31.0) have been
retrieved from the Jet Propulsion Laboratory's (JPL) horizons3 sys-
tem (Giorgini et al. 1996); this epoch is the t = 0 instant in the
simulations. Full details of these calculations can be found in de la
Fuente Marcos & de la Fuente Marcos (2012b, 2014, 2016a,b).
2.3 The case of Venus
Asteroid 2002 VE68 was confirmed as quasi-satellite of Venus by
Mikkola et al. (2004). The orbital evolution of this object was fur-
ther studied in de la Fuente Marcos & de la Fuente Marcos (2012b).
Its orbit is quite eccentric (e = 0.4103) and moderately inclined
(i = 9.◦0070). Fig. 1, top left-hand panel, shows the results of equa-
tions (1) from t = 0 until 10 yr later, i.e. over 16 orbital sidereal
periods of Venus. Over 16 analemmatic loops are displayed and,
consistent with its significant eccentricity, one of the lobes of the
analemma is very small, each loop resembling an inverted teardrop.
2.4 The case of the Earth
Our calculations show that our planet hosts the largest known pop-
ulation of quasi-satellites in the Solar system; however, their dy-
namical origin appears to be rather heterogeneous (de la Fuente
Marcos & de la Fuente Marcos 2014, 2016a,b). There are five con-
firmed quasi-satellites of the Earth: 164207 (2004 GU9) (Connors
et al. 2004; Mikkola et al. 2006; Wajer 2010), 277810 (2006 FV35)
(Wiegert et al. 2008; Wajer 2010), 2013 LX28 (Connors 2014),
2014 OL339 (de la Fuente Marcos & de la Fuente Marcos 2014,
2016a) and 469219 (2016 HO3)4 (de la Fuente Marcos & de la
Fuente Marcos 2016b). Fig. 1, top right-hand panel, shows the ap-
parent motion of these five objects over 10 sidereal periods; a wide
range of behaviours, from a very symmetric figure-eight to very
distorted teardrop shapes, is observed.
Asteroid 164207 (purple) has both moderate eccentricity (e =
0.1362) and inclination (i = 13.◦6491), and it traces a somewhat
symmetric figure-eight that slowly shifts, keeping the position of
1 Originally published in Russian, Astron. Zh., 50, 1071-1075 (1973
September -- October).
2 http://www.ast.cam.ac.uk/∼sverre/web/pages/nbody.htm
3 http://ssd.jpl.nasa.gov/?horizons
4 http://www.jpl.nasa.gov/news/news.php?feature=6537
MNRAS 000, 1 -- 6 (2016)
The analemma criterion for quasi-satellites
3
20
10
0
-10
-20
)
o
(
*
δ
-30
24
20
16
12
α* (o)
8
4
0
)
o
(
*
δ
40
20
0
-20
-40
)
o
(
δ
)
o
(
*
δ
80
60
40
20
0
-20
-40
-60
24
60
40
20
0
-20
-40
-60
-80
164207 (2004 GU9)
2014 OL339
469219 (2016 HO3)
2013 LX28
277810 (2006 FV35)
20
16
12
α (o)
8
4
0
2004 AE9
Comet 295P/LINEAR (2002 AR2)
241944 (2002 CU147)
8
4
0
12
α* (o)
24
20
16
12
α* (o)
8
4
0
24
20
16
)
o
(
*
δ
40
20
0
-20
-40
-60
-80
24
80
60
40
20
0
-20
-40
-60
)
o
(
*
δ
20
16
12
α* (o)
8
4
0
24
20
16
12
α* (o)
8
4
0
Figure 1. Apparent motion in host-centric equatorial coordinates of known quasi-satellites. Asteroid 2002 VE68 from Venus (top left-hand panel), 164207
(2004 GU9), 277810 (2006 FV35), 2013 LX28, 2014 OL339 and 469219 (2016 HO3) from the Earth (top right-hand panel), 76146 (2000 EU16) from Ceres
(middle left-hand panel), comet 295P/LINEAR (2002 AR2), 241944 (2002 CU147) and 2004 AE9 from Jupiter (middle right-hand panel), 63252 (2001 BL41)
from Saturn (bottom left-hand panel), and 309239 (2007 RW10) from Neptune (bottom right-hand panel).
the node almost fixed. Asteroid 277810 (gold) has significant ec-
centricity (e = 0.3776), but low orbital inclination (i = 7.◦1041);
consistently, its apparent motion describes a very distorted teardrop
as the effect of the eccentricity dominates. Asteroid 2013 LX28
(green) follows a quite eccentric (e = 0.4520) and very inclined
(i = 49.◦9754) path that translates into an apparent motion that
traces an elongated teardrop-shaped curve. Asteroid 2014 OL339
(blue) describes a somewhat elliptic analemma which suggests that
one of the effects is nearly negligible; consistently, it follows a very
eccentric (e = 0.4608) but moderately inclined (i = 10.◦1868) orbit.
Finally, the orbit pursued by 469219 (red) has both low eccentric-
ity (e = 0.1041) and inclination (i = 7.◦77140); consistently, the
analemma described in the sky resembles that of 164207 with a
rather symmetric shifting figure-eight.
2.5 The case of Ceres
Planets are not the only possible hosts of quasi-satellite bodies,
dwarf planet Ceres also has one of these interesting companions
(Christou 2000; Christou & Wiegert 2012), 76146 (2000 EU16).
Fig. 1, middle left-hand panel, shows the results of nearly 11
sidereal periods. Asteroid 76146 follows a low-eccentricity (e =
0.1674), low-inclination (i = 8.◦8475) orbit; surprisingly, its rela-
tively low eccentricity is high enough to induce a rather distorted
teardrop shape to the resulting analemma.
MNRAS 000, 1 -- 6 (2016)
2.6 The case of Jupiter
Jupiter is often regarded as the host of the largest known popu-
lation of quasi-satellites with at least six, including asteroids and
comets (Kinoshita & Nakai 2007; Wajer & Królikowska 2012).
However, we failed to confirm several of the proposed candidates
as present-day quasi-satellites of Jupiter. We have found three con-
firmed quasi-satellites of Jupiter: comet 295P/LINEAR (2002 AR2)
(Kinoshita & Nakai 2007; Wajer & Królikowska 2012), 241944
(2002 CU147) (Wajer & Królikowska 2012) and 2004 AE9 (Ki-
noshita & Nakai 2007; Wajer & Królikowska 2012). Fig. 1, mid-
dle right-hand panel, shows the results of 8.4 sidereal periods of
apparent motion for these objects. Comet 295P/LINEAR (red) fol-
lows a very eccentric orbit (e = 0.6460) that is only moderately
inclined (i = 14.◦6912); consistently, a rather asymmetric figure-
eight is observed. This object is unlikely to be the mysterious mi-
nor planet 7617 in van Houten et al. (1970), the angular elements
being too different. Asteroid 241944 (blue) pursues a relatively ec-
centric orbit (e = 0.3136) that is also quite inclined (i = 32.◦8906);
the relatively rapidly shifting analemma exhibits somewhat sym-
metric lobes. In sharp contrast, 2004 AE9 follows a very eccentric
(e = 0.6459) but nearly ecliptic (i = 1.◦6521) orbit and its apparent
motion as seen from Jupiter traces a very squashed teardrop. Again,
the angular elements of 2004 AE9 are very different from those of
the mysterious minor planet 7617.
4
C. de la Fuente Marcos and R. de la Fuente Marcos
2.7 The case of Saturn
Gallardo (2006) indicated that 15504 (1999 RG33) could be a quasi-
satellite of Saturn. Our calculations show that it is indeed a tran-
sient co-orbital of Saturn, but not a quasi-satellite like the objects
previously discussed. Its apparent motion somewhat resembles that
of Molniya-type artificial satellites of the Earth (the apocentre of
the very eccentric orbits occurs at a large declination). However,
63252 (2001 BL41) that was discovered by Gehrels et al. (2001)
is currently a short-lived quasi-satellite of Saturn that will change
its current dynamical status in about 130 yr from now. Asteroid
63252 follows an eccentric (e = 0.2948) but moderately inclined
(i = 12.◦5163) orbit. Fig. 1, bottom left-hand panel, shows the ap-
parent motion of this object as seen from Saturn for about 4.4 side-
real periods, prior to leaving its current quasi-satellite state. The
analemmatic loops described by this object are very distorted as a
result of its unstable dynamical behaviour and the lobes have some-
what different sizes because the effect derived from eccentricity is
stronger than that of inclination. Although its orbital evolution is
rather chaotic, 63252 -- an organic rich D-type asteroid (Dores-
soundiram et al. 2003) -- has been pre-selected by NASA for an
in situ exploration mission (Ryan et al. 2009).
2.8 The case of Neptune
Asteroid 309239 (2007 RW10) is so far the only confirmed quasi-
satellite of Neptune (de la Fuente Marcos & de la Fuente Marcos
2012a) and it is one of the largest known co-orbital companions
in the Solar system with a diameter of about 250 km. It follows
an eccentric (e = 0.3004) and rather inclined (i = 36.◦1755) orbit.
Fig. 1, bottom right-hand panel, shows 7.3 sidereal periods of a
very regular analemma of the figure-eight type with both lobes of
nearly the same size which confirms that the effects derived from
eccentricity and inclination have very similar strength in this case.
3 ACCIDENTAL QUASI-SATELLITES: THE CASE OF
PLUTINO 15810 (1994 JR1)
Quasi-satellites are not exclusive of planetary hosts as the case of
76146 (2000 EU16) and Ceres confirms. Yu & Tremaine (1999)
and Tiscareno & Malhotra (2009) used numerical simulations to
predict the existence of minor bodies experiencing quasi-satellite
behaviour with respect to dwarf planet Pluto. Plutino 15810 (1994
JR1) was identified as an accidental quasi-satellite of Pluto by de la
Fuente Marcos & de la Fuente Marcos (2012c). It was termed ac-
cidental because, for this object, λr circulates with a superimposed
libration resulting from the oscillation of the orbital period induced
by the 2:3 mean motion resonance with Neptune. Such libration
plays a role in triggering and terminating the quasi-satellite phase.
Porter et al. (2016) have used astrometry acquired by NASA's New
Horizons spacecraft to improve the already robust orbital solution
available for this object (see Appendix A) and revisit its quasi-
satellite status. The new data have been used to argue that the quasi-
satellite nature of 15810 must be rejected.5
Fig. 2 clearly shows that although the orbital solution of 15810
has been indeed greatly improved using New Horizons data (see
Appendix A), its orbital evolution still matches the one described
in de la Fuente Marcos & de la Fuente Marcos (2012c). In black,
5 http://www.nasa.gov/feature/new-horizons-collects-first-science-on-a-
post-pluto-object
)
o
(
*
δ
40
20
0
-20
-40
-60
-80
24
20
16
12
α* (o)
8
4
0
Figure 2. Apparent motion in Pluto-centric equatorial coordinates of 15810
(1994 JR1). The black curve shows the shifting analemma resulting from
the dynamical evolution of the latest version of the orbit that includes data
from NASA's New Horizons spacecraft (third orbital solution in Table A1).
The green curve is equivalent to the black one but using the second orbital
solution in Table A1.
we have the apparent motion resulting from the latest orbit avail-
able for 15810 (third orbital solution in Table A1). The figure
displays the time interval of interest -- the one showing analem-
matic behaviour -- that goes from 1200 years prior to t = 0 to 200
years afterwards or about 5.6 sidereal orbital periods of Pluto. The
analemma shifts rapidly and it is quite distorted because the orbits
of both Pluto and 15810 are eccentric and there is a chaotic inter-
action between the two bodies. Within the context of the analemma
criterion, the behaviour observed in Fig. 2 is not very different
from that of some of the objects in Fig. 1. The apparent motion of
15810 closely resembles that of comet 295P/LINEAR (2002 AR2)
or 63252 (2001 BL41). If the pre-NASA's New Horizons distant en-
counter orbit (second orbital solution in Table A1) is used (green
curve), the differences are minimal. Porter et al. (2016) argue that,
for 15810, instead of transient quasi-satellite behaviour we should
speak of periodic (every 2.4 Myr) scattering conjunctions; how-
ever, Fig. 2 suggests that 15810 is not different from comet 295P or
63252 in dynamical terms when the analemma criterion is applied
during one of its encounters with Pluto. Therefore, Pluto has at least
one present-day (transient and recurrent) quasi-satellite, 15810.
4 DISCUSSION
Our analysis so far argues in favour of a conjecture: objects that
follow a quasi-satellite path with respect to a host trace an analem-
matic curve in the sky as observed from the host over a sidereal or-
bital period. Conversely, any object tracing an analemmatic curve
in the sky as observed from the host over a sidereal year must be a
quasi-satellite of the host. Unfortunately, a rigorous mathematical
or even a numerical proof of the assumed theorem and its inverse
is out of the scope of this work because (1) a complete theory of
quasi-satellite motion still remains elusive (see e.g. Mikkola et al.
2006) and (2) the relevant volume of the orbital parameter space to
be explored is simply too large. Intuitively, the truthfulness of our
conjecture can hardly be argued. The position of the quasi-satellite
as seen from the host is subjected to two periodic librations. At
some point during the sidereal year, the quasi-satellite is east of
the host, very nearly half an orbital period later it is west from
the host. As the quasi-satellite bean-shaped loop (see e.g. fig. 1 in
Mikkola et al. 2004) drifts back and forth, the peri-host shifts from
eastwards to westwards from the host (this causes the loop drift).
This behaviour is mainly the result of the difference in eccentricity
between host and quasi-satellite and drives the oscillation in host-
centric right ascension. The relative inclination between the equa-
MNRAS 000, 1 -- 6 (2016)
torial plane of the host and the orbital plane of the quasi-satellite
drives the oscillation in host-centric declination. These two oscil-
lations have (nearly) commensurable frequencies because host and
quasi-satellite have very similar orbital periods and generate the
analemma.
Co-orbitals have been traditionally classified as such after the
statistical analysis of numerical simulations. However, an algo-
rithm to decide whether an object is a quasi-satellite of a given host,
not based on N-body simulations, is described in detail in section
4 of Mikkola et al. (2006). The analemmas or analemmata in Figs
1 and 2 show that, in the particular case of quasi-satellites, astrom-
etry can be readily used to perform a reliable classification. Figs
1 and 2 also show that both very regular -- 164207 (2004 GU9),
469219 (2016 HO3) or 309239 (2007 RW10) -- and rather irregular
-- comet 295P/LINEAR (2002 AR2), 63252 (2001 BL41) or 15810
(1994 JR1) -- short-term evolutions are possible. The analemma
traced by the quasi-satellite encodes relevant orbital information.
Distorted, rapidly shifting analemmas are characteristic of quasi-
satellites moving in strongly perturbed orbits.
Our calculations show that, if a suitable quasi-satellite is
found, it can be used as a permanent platform to install instru-
mentation that may be used to monitor permanently the host body
and enable a relatively stable communications relay for subsequent
missions (e. g. landing quasi-autonomous vehicles on the host) at
zero fuel cost because the quasi-satellite behaves like a geosyn-
chronous satellite from the point of view of the host body. For this
task, the smaller its average distance from the host the better (see
e.g. the case of 469219 as discussed in de la Fuente Marcos & de la
Fuente Marcos 2016b). Artificial quasi-satellites are also possible
(see e.g. Kogan 1989 for the Phobos mission). For quasi-satellites
sufficiently close to a host, substantial parallax may occur; there-
fore, and depending on the location of the observer on the surface
of the host, different analemmas may be observed. This issue to-
gether with the shift of the analemma loop induced by orbital evo-
lution requires robotic tracking of the quasi-satellite yearly move-
ment around its analemma. The use of quasi-satellite trajectories in
astrodynamics has been frequently discussed (see e.g. Kogan 1989;
Lidov & Vashkov'yak 1993, 1994; Mikkola & Prioroc 2016). It
could be the case that the analemmatic behaviour described here
had been found before within the context of astrodynamical studies
(see e.g. Kogan 1989).
5 CONCLUSIONS
In this paper, we have explored a new criterion to identify quasi-
satellites. In sharp contrast with the numerical strategies customar-
ily applied in the study of co-orbital bodies, the criterion described
here can make direct use of observational astrometric data. Our
conclusions can be summarized as follows.
(i) Bona fide quasi-satellites trace paths in the sky which repeat
every sidereal period when observed from their hosts. These
paths can be described as analemmatic curves similar to those
found for geosynchronous orbits. The analemma shifts as the
orbit of the quasi-satellite changes over time.
(ii) The existence of this analemmatic behaviour turns quasi-
satellites, natural or artificial, into potentially interesting plat-
forms for the future of space exploration.
(iii) The Earth has the largest known number of present-day
quasi-satellites, five. Jupiter comes in second place with three.
Venus, Saturn, Neptune and dwarf planet Ceres have one
each.
MNRAS 000, 1 -- 6 (2016)
The analemma criterion for quasi-satellites
5
(iv) Applying the analemma criterion, Plutino 15810 (1994 JR1)
is as good a quasi-satellite as it may get. Therefore, dwarf
planet Pluto hosts at least one quasi-satellite at present.
(v) Asteroid 63252 (2001 BL41) is a present-day transient quasi-
satellite of Saturn.
(vi) Historically, the first object identified as quasi-satellite (in
this case of Jupiter) was an asteroid moving in a comet-like
orbit. Unfortunately, this object appears to have been lost
since its announcement back in 1973.
ACKNOWLEDGEMENTS
We thank the referee, S. Mikkola, for his prompt reports, S. J.
Aarseth for providing the code used in this research, and S. B.
Porter for discussing the results of his group with us prior to pub-
lication. In preparation of this paper, we made use of the NASA
Astrophysics Data System, the ASTRO-PH e-print server, and the
MPC data server.
REFERENCES
Aarseth S. J., 2003, Gravitational N-body simulations. Cambridge Univ.
Press, Cambridge, p. 27
Chalmers J. S., 1987, Am. J. Phys., 55, 548
Chebotarev G. A., 1974, SvA, 17, 677
Christou A. A., 2000, A&A, 356, L71
Christou A. A., Wiegert P., 2012, Icarus, 217, 27
Connors M., 2014, MNRAS, 437, L85
Connors M., Veillet C., Brasser R., Wiegert P., Chodas P., Mikkola S., In-
nanen K., 2004, Meteoritics Planet. Sci., 39, 1251
de la Fuente Marcos C., de la Fuente Marcos R., 2012a, A&A, 545, L9
de la Fuente Marcos C., de la Fuente Marcos R., 2012b, MNRAS, 427, 728
de la Fuente Marcos C., de la Fuente Marcos R., 2012c, MNRAS, 427, L85
de la Fuente Marcos C., de la Fuente Marcos R., 2014, MNRAS, 445, 2961
de la Fuente Marcos C., de la Fuente Marcos R., 2016a, MNRAS, 455, 4030
de la Fuente Marcos C., de la Fuente Marcos R., 2016b, MNRAS, in press
(arXiv:1608.01518)
di Cicco D., 1979, Sky Telescope, 57, 536
Doressoundiram A., Tozzi G. P., Barucci M. A., Boehnhardt H., Fornasier
S., Romon J., 2003, AJ, 125, 2721
Gallardo T., 2006, Icarus, 184, 29
Gehrels T., Gleason A. E., McMillan R. S., Montani J. L., Larsen J. A.,
Marsden B. G., 2001, MPEC Circ., MPEC 2001-B44
Giorgini J. D. et al., 1996, BAAS, 28, 1158
Heath W., 1923, The Observatory, 46, 286
Holbrow C. H., 2013, e-print arXiv:1302.0765
Irvine S., 2001, J. R. Astron. Soc. Can., 95, 273
Jackson J., 1913, MNRAS, 74, 62
Kinoshita H., Nakai H., 2007, Celest. Mech. Dyn. Astron., 98, 181
Kogan A. Y., 1989, Cosmic Res., 26, 705
Lidov M. L., Vashkov'yak M. A., 1993, Cosmic Res., 31, 187
Lidov M. L., Vashkov'yak M. A., 1994, Astron. Lett., 20, 188
Makino J., 1991, ApJ, 369, 200
Mikkola S., Brasser R., Wiegert P., Innanen K., 2004, MNRAS, 351, L63
Mikkola S., Innanen K., Wiegert P., Connors M., Brasser R., 2006, MN-
RAS, 369, 15
Mikkola S., Prioroc C.-L., 2016, MNRAS, 457, 1137
Morais M. H. M., Morbidelli A., 2002, Icarus, 160, 1
Murray C. D., Dermott S. F., 1999, Solar System Dynamics, Cambridge
Univ. Press, Cambridge, p. 97
Oliver B. M., 1972, Sky Telescope, 44, 20
Porter S. B. et al., 2016, ApJL, submitted (arXiv:1605.05376)
Raisz E., 1942, Sky Telescope, 1, 11
Ryan E. L. et al., 2009, Am. Astron. Soc. -- DPS meeting, 41, 16.26
6
C. de la Fuente Marcos and R. de la Fuente Marcos
Tiscareno M. S., Malhotra R., 2009, AJ, 138, 827
van Houten C. J., van Houten-Groeneveld I., Herget P., Gehrels T., 1970,
A&AS, 2, 339
Wajer P., 2010, Icarus, 209, 488
Wajer P., Królikowska M., 2012, Acta Astron., 62, 113
Wiegert P. A., DeBoer R., Brasser R., Connors M., 2008, J. R. Astron. Soc.
Can., 102, 52
Yu Q., Tremaine S., 1999, AJ, 118, 1873
APPENDIX A: COMPARISON BETWEEN THE PRE- AND
POST-NEW HORIZONS DATA ORBITAL SOLUTIONS
Table A1 shows three orbital solutions for 15810 (1994 JR1). The
third column corresponds to the one currently available and in-
cludes astrometry acquired by NASA's New Horizons spacecraft
and discussed in Porter et al. (2016). As pointed out by Porter et
al. (2016), the New Horizons data have improved the orbital solu-
tion of 15810 very significantly. However, numerical simulations
equivalent to those in de la Fuente Marcos & de la Fuente Marcos
(2012c) but making use of the second and third orbital solutions in
Table A1 still produce the same basic results; the differences are
simply too small to claim any dramatic change in the nature of the
orbital evolution of 15810 as a result of the new and indeed im-
proved orbit. The original description in de la Fuente Marcos &
de la Fuente Marcos (2012c) is certainly still valid. Nevertheless,
the astrometry discussed in Porter et al. (2016) confirms beyond
any doubt the role that NASA's New Horizons spacecraft may play
in improving the orbital solutions of many trans-Neptunian objects
over the next decade or so.
This paper has been typeset from a TEX/LATEX file prepared by the author.
MNRAS 000, 1 -- 6 (2016)
The analemma criterion for quasi-satellites
7
Table A1. Heliocentric Keplerian orbital elements of 15810 (1994 JR1) from JPL's Small-Body Database and horizons On-Line Ephemeris System; values
include the 1σ uncertainty. The orbit in the left-hand column was the one available back in 2012 and it was used by de la Fuente Marcos & de la Fuente Marcos
(2012c); this orbit is referred to the epoch 2456200.5 JD CT (2012-September-30.0) and it was computed using 43 observations with an arc-length of 2236 d.
The orbital solution in the column next to it was computed on 2015 October 05 13:55:21 ut and it is referred to the epoch 2457600.5 JD TDB (2016-July-31.0)
TDB; it was computed using 49 observations with an arc-length of 7701 d. The third orbit is the one currently available and it was computed on 2016 June 21
15:49:21 ut. This new and improved orbital solution includes astrometry acquired by NASA's New Horizons spacecraft and is referred to the same epoch as
the previous one; it was computed using 78 observations with an arc-length of 8002 d.
Semimajor axis, a (au)
Eccentricity, e
Inclination, i (◦)
Longitude of the ascending node, Ω (◦)
Argument of perihelion, ω (◦)
=
=
=
=
=
39.24±0.02
0.1143±0.0003
3.8032±0.0002
144.753±0.011
102.1±0.2
39.427±0.011
0.1196±0.0002
3.80801±0.00005
144.6859±0.0011
101.55±0.03
39.4224±0.0009
0.119501±0.000010
3.80802±0.00005
144.6854±0.0007
101.535±0.012
MNRAS 000, 1 -- 6 (2016)
|
1607.03602 | 1 | 1607 | 2016-07-13T06:26:13 | Non-thermal production and escape of OH from the upper atmosphere of Mars | [
"astro-ph.EP",
"physics.chem-ph"
] | We present a theoretical analysis of formation and kinetics of hot OH molecules in the upper atmosphere of Mars produced in reactions of thermal molecular hydrogen and energetic oxygen atoms. Two major sources of energetic O considered are the photochemical production, via dissociative recombination of O$_{2}^{+}$ ions, and energizing collisions with fast atoms produced by the precipitating Solar Wind (SW) ions, mostly H$^+$ and He$^{2+}$, and energetic neutral atoms (ENAs) originating in the charge-exchange collisions between the SW ions and atmospheric gases. Energizing collisions of O with atmospheric secondary hot atoms, induced by precipitating SW ions and ENAs, are also included in our consideration. The non-thermal reaction O + H$_2(v,j) \rightarrow$ H + OH$(v',j')$ is described using recent quantum-mechanical state-to-state cross sections, which allow us to predict non-equilibrium distributions of excited rotational and vibrational states $(v',j')$ of OH and expected emission spectra. A fraction of produced translationally hot OH is sufficiently energetic to overcome Mars' gravitational potential and escape into space, contributing to the hot corona. We estimate the total escape flux from dayside of Mars for low solar activity conditions at about $5\times10^{22}$ s$^{-1}$, or about 0.1\% of the total escape rate of atomic O and H. The described non-thermal OH production mechanism is general and expected to contribute to the evolution of atmospheres of the planets, satellites, and exoplanets with similar atmospheric compositions. | astro-ph.EP | astro-ph |
Non-thermal production and escape of OH from the
upper atmosphere of Mars
M. Gacesaa, N. Lewkowb, V. Kharchenkoc,b
aNASA Ames Research Center, Moffett Field, CA 94035
bDepartment of Physics, University of Connecticut, Storrs, CT 06268
cInstitute for Theoretical Atomic and Molecular Physics, Harvard-Smithsonian Center
for Astrophysics, Cambridge, MA 02138
Abstract
We present a theoretical analysis of formation and kinetics of hot OH molecules
in the upper atmosphere of Mars produced in reactions of thermal molecu-
lar hydrogen and energetic oxygen atoms. Two major sources of energetic
O considered are the photochemical production, via dissociative recombina-
tion of O+
2 ions, and energizing collisions with fast atoms produced by the
precipitating Solar Wind (SW) ions, mostly H+ and He2+, and energetic neu-
tral atoms (ENAs) originating in the charge-exchange collisions between the
SW ions and atmospheric gases. Energizing collisions of O with atmospheric
secondary hot atoms, induced by precipitating SW ions and ENAs, are also
included in our consideration. The non-thermal reaction O + H2(v, j) → H +
OH(v(cid:48), j(cid:48)) is described using recent quantum-mechanical state-to-state cross
sections, which allow us to predict non-equilibrium distributions of excited
rotational and vibrational states (v(cid:48), j(cid:48)) of OH and expected emission spec-
tra. A fraction of produced translationally hot OH is sufficiently energetic to
overcome Mars' gravitational potential and escape into space, contributing
to the hot corona. We estimate the total escape flux from dayside of Mars for
low solar activity conditions at about 5× 1022 s−1, or about 0.1% of the total
escape rate of atomic O and H. The described non-thermal OH production
mechanism is general and expected to contribute to the evolution of atmo-
spheres of the planets, satellites, and exoplanets with similar atmospheric
compositions.
Email addresses: [email protected] (M. Gacesa), [email protected]
(N. Lewkow), [email protected] (V. Kharchenko)
Preprint submitted to Icarus
October 11, 2018
Keywords:
1. Introduction
The escape of volatile atmospheric species to space is important for un-
derstanding the evolution of Mars' atmosphere and climate as it transi-
tioned from the conditions that supported liquid water into the cold, dry,
low-pressure climate that we witness today (McElroy and Donahue, 1972;
Johnson et al., 2008; Lammer et al., 2013; Lillis et al., 2015). While it is
well established that the evaporation of the martian atmosphere is driven by
the interaction with the solar radiation and interplanetary plasma, with the
absence of an intrinsic planetary magnetic field and Mars' lower mass ac-
celerating the process (Chassefi`ere and Leblanc, 2004; Johnson et al., 2008),
detailed physical mechanisms and their mutual interactions are still not fully
understood and 3D global atmospheric models cannot simultaneously ex-
plain all observed effects (Lillis et al., 2015; Lee et al., 2015a,b). Attempting
to resolve the remaining unanswered questions and shed light on water in-
ventory in the early history of Mars is the main scientific objective of the
ongoing NASA's Mars Atmosphere and Volatile Evolution (MAVEN) mission
(Jakosky et al., 2015; Bougher et al., 2015).
At the present time, the atmospheric escape from Mars is comprised
of thermal (Jeans) escape and various non-thermal mechanisms, including
photo-chemical escape of neutrals (Chassefi`ere and Leblanc, 2004; Johnson
et al., 2008; Lammer et al., 2013; Lee et al., 2015b) and escape of ions gov-
erned by the interplay of the solar wind with the induced martian magne-
tosphere and crustal magnetic fields (Acuna et al., 1999; Nagy et al., 2004;
Dong et al., 2015; Rahmati et al., 2015). Major escaping species include
atomic hydrogen, oxygen, and carbon, of which the first two directly affect
the estimates of water abundance on primordial Mars (Lammer et al., 2013).
A major photochemical process responsible for escape of neutrals heavier
than hydrogen is dissociative recombination (DR) of O+
2 , which serves as a
major source of hot O atoms that either directly escape to space or form a hot
oxygen corona (Ip, 1988; Fox, 1993; Krest'yanikova and Shematovich, 2005;
Lee et al., 2015b; Deighan et al., 2015). The nascent suprathermal O atoms
can collide with thermal background gases in the upper atmosphere and
transfer sufficient kinetic energy to eject them to space. This non-thermal
escape mechanism of light elements, also known as a collisional ejection,
2
was studied for He atoms (Bovino et al., 2011), and H2 and HD molecules
(Gacesa et al., 2012), which were found to produce significant fluxes and
possibly affect the H/D ratio in Mars' upper atmosphere by 5-10%.
One of the major goals of our research, reported in this article, is to de-
velop a consistent model of production of non-thermal atoms and molecules in
the Martian atmosphere and describe non-equilibrium atmospheric reactions
caused by hot particles. In this study, we explore reactive collisions of hot O
atoms with H2 molecules, leading to formation of rotationally-vibrationally
(RV) excited OH molecules in Mars' upper atmosphere. The description
of the reaction is based on quantum-mechanical state-to-state reactive cross
sections at high temperatures (Gacesa and Kharchenko, 2014), while kinetic
theory is used to calculate the energy transfer to translational and internal
degrees of freedom of the products. We use a 1D model of the martian atmo-
sphere to estimate altitude profiles of the total formation and escape rates
of OH molecules and non-thermal RV distributions for a selected orbital ge-
ometry and solar activity. In addition to the DR, secondary hot O atoms
energized in collisions with energetic neutral atoms (ENAs) (Lewkow and
Kharchenko, 2014) are considered as an efficient source of hot O atoms in
our model.
2. Methods
2.1. Hot atom sources
We consider dissociative recombination (DR) of oxygen molecular ions
2 + e → O + O + ∆E, as the main source of hot O
with electrons, O+
atoms. This process is widely accepted as the dominant source of O in the
martian ionosphere in the current epoch (Ip, 1988; Fox, 1993; Krest'yanikova
and Shematovich, 2005). The DR of O+
2 is an exothermic process that can
proceed via five dissociation pathways and produce energetic O(3P), O(1D),
and O(1S) with excess kinetic energy between 0.84 eV and 6.99 eV per a pair
(Guberman, 1988; Fox and Ha´c, 2009). Since the state-to-state cross sections
are available only for the reaction of O(3P) + H2, we assume that all reactive
collisions involve hot O(3P). This approximation is further justified by the
fact that the lifetime of energetic metastable O(1D) atoms is short before
they relax into their ground state via spontaneous emission or collisional
quenching with atmospheric gases (Kharchenko et al., 2005).
To evaluate the contribution of the photo-chemical production mecha-
nism of energetic O atoms, we followed the approach of Bovino et al. (2011)
3
Figure 1: Energy distributions of hot O(3P ) produced by DR of O+
2 and in collisions
with energetic neutral H and He atoms. The distributions are normalized to unity and
calculated at 200 km.
to construct the energy distribution and rate of production of hot O atoms,
fDR(Ei, h), as a function of the altitude h and the initial energy Ei of the
O atoms dissociated via i-th channel. Four significant dissociation channels
were considered with branching ratios and exothermicities given by Guber-
man (1988) and Fox and Ha´c (2009). We constructed fDR(Ei, h) distributions
of nascent O atoms below 400 km for conditions of low solar activity (Fox
and Ha´c, 2009) and adopted the atmospheric model by Krasnopolsky (2010)
with extrapolated hot O source functions from Fox and Ha´c (2009).
As a secondary source of hot O atoms we considered elastic collisions
of precipitating ENAs, namely energetic H and He, with thermal atmo-
spheric oxygen, leading to production of secondary hot O atoms (SHOAs)
(Krest'yanikova and Shematovich, 2005; Wang et al., 2013; Lewkow and
Kharchenko, 2014). Incoming ENAs decelerate in the martian atmosphere
by transferring a fraction of its kinetic energy to thermal atmospheric gases
4
012345Energy (eV)00.511.52Probability Density (1/eV)DR of O2+He ENAH ENAtypically in thousands of collisions dominated by small scattering angles. The
SHOAs are produced mainly at altitudes lower than 250 km, where the atmo-
spheric density sharply increases, and are predicted to have kinetic energies
up to 4 eV (Lewkow and Kharchenko, 2014). We carried out a 3D Monte
Carlo simulation of thermalization of ENAs in the martian atmosphere for
average solar wind activity and constructed altitude-dependent SHOA source
function fENA(h). A distinct feature of our MC approach are energy transfer
cross sections given as functions of both collision velocity and angle, resulting
in strongly forward-peaked anisotropic distributions. The simulation is de-
scribed in detail in Lewkow and Kharchenko (2014). In this work we adopted
similar physical parameters, including the ensemble sizes, as in the article.
The energy distributions (normalized to unity) of hot O and its volume
production rates as a function of altitude are given for DR and ENA produc-
tion mechanisms in Fig. 1 and Fig. 2, respectively. The total source function
f (Ei, h), given in the units of volume production rate, was constructed as
a sum of distributions produced by considering both non-thermal processes,
f (Ei, h) = fDR(Ei, h) + fENA(Ei, h). For the altitudes higher than 250 km
and lower than 130 km (in case of DR of O+
2 ), the production rate distribu-
tions were smoothly extrapolated with descending exponential functions to
avoid unphysical discontinuities. The final results were not found to be de-
pend much on the exact functional forms due to the fact that the production
rates in the extrapolated regions are very small. The energy distributions
given in Fig. 1 were calculated for the altitude of 200 km and do not change
significantly for the altitudes at which majority of thermalization processes
take place (Zhang et al., 2009; Fox and Ha´c, 2009).
2.2. Cross sections and kinetics of non-thermal OH production
We use quantum-mechanical state-to-state reactive cross sections and dif-
ferential cross sections (DCSs) (Gacesa and Kharchenko, 2014) to model the
chemical reaction O(3P)+H2(v, j) → H + OH(v(cid:48), j(cid:48)), where (v, j) and (v(cid:48), j(cid:48))
are the initial and final RV levels of H2 and OH, respectively. The cross sec-
tions of Gacesa et al. (2012) are adopted for elastic scattering. Both sets of
cross sections were constructed in an extended collision energy range for pur-
poses of modeling non-thermal atmospheric processes and high-temperature
combustion.
Momentum transfer reactive cross sections are evaluated according to
5
Figure 2: Altitude profile of the rate of production of secondary hot O(3P ) atoms by DR
of O+
2 and collisions of precipitating H and He ENAs with atmospheric O. Mean solar
activity is assumed.
(Parker and Pack, 1978)
σmt
vj,v(cid:48)j(cid:48)(E) = 2π
(cid:90) ∞
0
Qvj,v(cid:48)j(cid:48)(E, θ)
(cid:32)
1 − cos θ
(cid:33)
(cid:114)
E(cid:48)
E
sin θdθ,
(1)
where Qvj,v(cid:48)j(cid:48)(E, θ) are reactive differential cross sections (DCSs) and E(cid:48) =
E − (Ev(cid:48)j(cid:48) − Evj) is the final energy of the product OH. Here, the internal RV
energies of H2(v, j) and OH(v(cid:48), j(cid:48)) are given by Evj and Ev(cid:48)j(cid:48), respectively.
The RV energies Evj and Ev(cid:48)j(cid:48) were found using a mapped Fourier Grid
Method (Kokoouline et al., 1999) in the asymptotic limits (corresponding to
diatomic H2 and OH) of the potential energy surfaces used in Gacesa and
Kharchenko (2014).
The kinetic energy transfer rate from hot O atoms to the product OH(v(cid:48), j(cid:48))
is determined using kinetic theory with quantized internal molecular degrees
6
10-310-210-1100101102103Production rate of hot O atoms by DR and ENAs (cm-3s-1)100200Altitude (km)H (ENA)He (ENA)O (DR)Figure 3: Fraction Γv(cid:48)j(cid:48)
vj (E) of the OH(v(cid:48), j(cid:48)) molecules produced in the reaction O+H2(v =
0, j = 0) having sufficient kinetic energy to escape from Mars given as a function of O atom
kinetic energy. Selected rotationally (left panel) and vibrationally (right panel) excited
states of OH are shown.
of freedom and anisotropic cross sections as (Johnson, 1982)
(cid:0)1 + γv(cid:48)j(cid:48) − 2
√
γv(cid:48)j(cid:48) cos θ(cid:1) E,
(2)
Tv(cid:48)j(cid:48) =
mO mOH
(mO + mOH)2
where mO and mOH are masses of O and OH, respectively, E is the collision
energy in the laboratory frame (LF), γv(cid:48)j(cid:48) = v(cid:48)j(cid:48)/, is the ratio of and v(cid:48)j(cid:48),
the translational kinetic energies before and after the reaction in the center-
of-mass (CM) frame, respectively, and θ is the scattering angle in the CM
frame.
Following the reaction, the fraction of OH molecules sufficiently energetic
to escape is given by
Γv(cid:48)j(cid:48)
vj (E) =
(cid:82) π
(cid:82) π
Qvj,v(cid:48)j(cid:48)(E, θ) (1 − cos θ) sin θdθ
0 Qvj,v(cid:48)j(cid:48)(E, θ) (1 − cos θ) sin θdθ
θmin
,
(3)
where the critical angle θmin was determined by solving Eq.
(2) for the
angle, while requiring that the translational part of the transferred kinetic
7
234Energy of O atom (eV)00.20.40.60.81Γvjv'j'(E)j'=010203035234Energy of O atom (eV)v'=0123456energy, Tv(cid:48),j(cid:48) is equal to the escape energy of OH from Mars, EOH
esc = 2.08
eV. The resulting fractions Γv(cid:48)j(cid:48)
vj (E) for the vibrational states up to vmax = 6
and jmax = 36 were calculated and illustrated in Fig. 3 for selected values
of (v(cid:48), j(cid:48)) states. According to the kinetic model, the OH molecules could
be produced in the reaction in higher RV levels but they would not receive
sufficient kinetic energy to be able to escape.
2.3. 1D transport model
To determine the transport of non-thermal atoms and molecules, includ-
ing altitude profiles of OH, we use a 1D model built upon analogous as-
sumptions used in the study of escape of neutral He atoms and H2 molecules
from Mars (Bovino et al., 2011; Gacesa et al., 2012). In our model, we use
altitude-dependent rates of production of hot oxygen, f (E, h) by both DR of
O+
2 and precipitating ENAs, as described in the previous section. The atten-
uation of fluxes of hot O and OH is estimated based on their mean free path
in thermal atmospheric gases. Such an approach is a compromise between
a simple exobase approximation and more complex Monte Carlo simulation.
In comparison, our multi-species 3D Monte Carlo simulation of the ENA-
induced escape of hydrogen predicted about 40% higher escape rates than a
1D model (Lewkow and Kharchenko, 2014).
With the above assumptions, the volume production rate of hot OH(v(cid:48), j(cid:48))
at the altitude h can be written as
Pv(cid:48)j(cid:48)(h) =
1
2
×
(cid:90) ∞
(cid:90) h
0
hmin
2
dE TOH(h, E)nH2(h)Γv(cid:48)j(cid:48)
vj (E)σmt
vj,v(cid:48)j(cid:48)(E)
dh2f (E, h2)TO(h2, h, E),
(4)
where f (E, h) is the total hot O source function, Γv(cid:48)j(cid:48)
vj (E) is the fraction of the
OH molecules energetically allowed to escape, σmt
vj,v(cid:48)j(cid:48)(E) is the momentum
transfer reactive cross section, and nH2(h) is the density of atmospheric H2
gas taking part in reactions. The prefactor 1/2 implies that one half of
the produced energetic OH molecules are scattered towards the planet and
cannot escape regardless of the energy transferred. The transparencies TOH
and TO are defined as
(cid:35)
TOH(h, E) = exp
OH,i (E)ni(h(cid:48))
σel,mt
(cid:34)
−
(cid:90) hmax
dh(cid:48)(cid:88)
i
h
8
TO(h2, h, E) = exp
(cid:34)
−
(cid:90) h
h2
dh(cid:48)(cid:88)
i
(cid:35)
O,i (E)ni(h(cid:48))
σel,mt
,
(5)
and describe the loss of flux due to the collisions with thermal atmospheric
species. Specifically, TOH is equal to the escape probability of hot OH(v(cid:48), j(cid:48))
produced in collisions with the incident hot O of energy E at the altitude h,
while the TO is defined as the probability that the hot O atoms, produced
at the altitude h2, reach the altitude h without the energy loss in collisions
with other atmospheric constituents. The sums describe the loss of O and OH
flux in collisions with major constituents of the martian upper atmosphere,
where, for the i-th atmospheric species, ni(h) is the density, and σel,mt
H2,i (E) and
σel,mt
O,i (E) are elastic momentum transfer cross sections. Eight species were
included in the summation: CO2, CO, N2, O2, H2, H, Ar, and H. We use
only the elastic cross sections, which are about an order of magnitude larger
than reactive or inelastic cross sections, to evaluate the transparencies. The
integrals were evaluated using hmin
and hmax as practical integration limits,
taken to be 130 km and 800 km, respectively.
We used the momentum transfer cross sections for OH(v(cid:48), j(cid:48))+H scatter-
ing calculated in our previous work (Gacesa and Kharchenko, 2014). For
collisions with the other species we approximated the cross sections by mass-
scaling from known species: OH-He (from Ar-H2), OH-N2, OH-O2, OH-CO
(from O-H2), and OH-CO2 (from O-N2 (Balakrishnan et al., 1998)). Simi-
larly, the cross sections for O-N2, O-O2, O-He, and O-H were adopted from
Balakrishnan et al. (1998), Brunetti et al. (1981), Bovino et al. (2011), and
Zhang et al. (2009), respectively. As an additional simplification, we did not
include O-O collisions in the flux attenuation since they do not reduce the
total kinetic energy in the upward flux of hot O atoms. We also assume
that all reactive collisions of O+H2 take place with molecular hydrogen in
its ground state, (v = 0, j = 0), while at T = 280 K, first three rotational
states (j = 0, 1, 2) will be significantly populated. We do not expect this
approximation to introduce significant errors since the differences in reactive
cross sections between these states are smaller than 3%.
2
3. Altitude profile of hot OH
Using Eq. (4) we have calculated volume production rate of hot OH(v(cid:48), j(cid:48))
molecules in the reaction H2(v = 0, j = 0) + O at the altitudes ranging from
hmin = 130 km to hmax = 700 km, for all energetically-allowed target RV
9
Figure 4: Altitude profiles of volume production rates of vibrationally excited OH(v(cid:48) =
0 . . . 6) produced in reactions with non-thermal O atoms. The rates are summed over
rotational states j(cid:48) = 0 . . . 36 of OH(v(cid:48), j(cid:48)).
states, v(cid:48) = 0 . . . 6 and j(cid:48) = 0 . . . 36. We have included both sources of hot O,
DR and collisions with H and He ions and atoms, and performed the calcula-
tion with and without (by setting the escaping fraction factor Γv(cid:48)j(cid:48)
vj (E) = 1)
the constraint on the kinetic energy of the product OH. The resulting al-
j(cid:48) Pv(cid:48)j(cid:48)(h), are
shown in Figs. 4 and 5, for the total and escaping volume production rate,
respectively.
titude profiles of vibrationally excited hot OH, Pv(cid:48)(h) = (cid:80)
The volume production rates of hydroxyl molecules are very dependent
on density of the atmosphere and availability of hot oxygen. They peak at
about 210-220 km and remain significant below the exobase, up to about
170 km, below which the martian atmosphere becomes sufficiently dense to
allow more efficient diffusive mixing and fast thermalization of hot oxygen.
The production rates due to the ENA sources of hot O are about 500 times
smaller than due to the DR, which is understood to be the dominant process
10
10-610-510-410-310-2Volume production rate of OH molecules (cm-3s-1)200300400500600700Altitude (km)TOTALv' = 0v' = 1v' = 2v' = 3v' = 4v' = 5v' = 6Figure 5: Altitude profiles of volume production rates of vibrationally excited nascent
OH(v(cid:48) = 0 . . . 6) sufficiently energetic to escape produced in reactions with non-thermal O
atoms. The rates are summed over rotational states j(cid:48) = 0 . . . 36 of OH(v(cid:48), j(cid:48)) and shown
for both sources of hot oxygen (thin lines for ENA).
of non-thermal escape. Our model predicts that the non-thermal reaction
will mostly produce OH in v(cid:48) = 0 vibrational state, followed by higher vibra-
tional states up to v(cid:48) = 6, which comprises less than 1% of the population.
A similar range of excited vibrational states is predicted for the OH ener-
getically capable of escaping from Mars, even though the ratios between the
individual excited states are somewhat different. This is particularly the case
in enhanced rates for escaping OH(v(cid:48) = 2), which are comparable to those
for OH(v(cid:48) = 1) state.
We calculated the total production rates and the production rates of
OH(v(cid:48), j(cid:48)) capable of escaping as the product of the surface area of the dayside
of Mars and the non-thermal flux φv(cid:48)j(cid:48) given by:
φv(cid:48)j(cid:48) =
Pv(cid:48)j(cid:48)(h)dh
(6)
(cid:90) hmax
hmin
11
10-910-810-710-610-510-410-310-2Volume production rate of escaping OH molecules (cm-3s-1)200300400500600700Altitude (km)TOTALv' = 0v' = 1v' = 2v' = 3v' = 4v' = 5v' = 6DR + ENAENA onlyFigure 6: Top: Total production rates of OH(v(cid:48), j(cid:48)) as a function of its rotational and
vibrational states. Bottom: As above, for escaping OH(v(cid:48), j(cid:48)).
The total production rates of OH(v(cid:48), j(cid:48)) are shown in Fig. 6, where we also
show distributions of rotational states j(cid:48) for each vibrational state v(cid:48). In case
of the OH molecules sufficiently energetic to escape, the rates correspond to
the number of OH molecules in a state (v(cid:48), j(cid:48)) potentially escaping from Mars
at hmax = 700 km and can be compared with other escape processes. On
the other hand, the total production rate, unconstrained with respect to the
kinetic energy, gives an indication of the total number of excited OH(v(cid:48), j(cid:48))
molecules distributed according to the altitude profiles given in Fig. 4 on
the dayside of the planet and can be used to evaluate the brightness of the
emission spectra and Meinel bands of OH (Meinel, 1950). The total escape
rates of OH, summed over all internal states and compared with the rates
for Jeans and non-thermal escape of O, H, and H2 from literature, are given
in Table 1.
The produced hydroxyl molecule will decay mostly by predissociation,
followed by direct photo-destruction, while interactions with solar wind ions
are not expected to play a major role. The predissociation rates vary up
12
101910201021Prod. rate of OH (s-1)0102030Rotational state j'101910201021Prod. rate of OH (s-1)v' = 0v' = 1v' = 2v' = 3v' = 4v' = 5v' = 6Total rate (Γ = 1)Escaping OH onlyJeans (×1023 s−1)
NT (×1023 s−1)
OH H2
5.8a
1.0a
0.53
O
H
530b
170-410c
Table 1: Total non-thermal (NT) escape rates of OH from the ideal dayside of Mars
calculated at 700 km, assuming Mars' volumetric mean radius equal to 3389.5 km and an
ideal dayside. Published escape rates for H2, O, and H are shown for comparison (aGacesa
and Kharchenko (2014), bJohnson et al. (2008), cLee et al. (2015b)).
to about 30% with solar activity, and give a range of OH lifetimes of (1.13-
1.68)×105 sec for solar minimum conditions, considered in our calculation,
and (1.04 − 1.49) × 105 sec for the solar maximum, with uncertainties of up
to 20% (Budzien et al., 1994). The most important processes responsible for
destruction of OH in the lower atmosphere of Mars include reactions with
O, N, C, and H2, with the reaction rates of the order of (1 − 7) × 10−10
cm3s−1. The collisions with O will be a dominant mechanism of destruction
at lower altitudes, where we can estimate the lifetime of OH to be of the
order of 10-100 seconds, depending mainly on the atomic oxygen densities.
This does not limit the escaping flux significantly. The lifetime of excited OH
in the v(cid:48) = 2 and v(cid:48) = 1 is about 12 and 24 ms (van Dishoeck and Dalgarno,
1984), respectively, which allows the OH formed in the upper atmosphere to
travel between 60-75 km (v = 1) or 115-150 km (v = 2) before decaying,
mainly by spontaneous emission. Higher vibrational states will decay faster,
resulting in faint Meinel emission bands localized at the regions where the
OH production rates are the largest.
4. Conclusions
We report the first theoretical model of non-thermal formation of excited
OH molecules in the upper atmosphere of Mars in reactions of translation-
ally hot O atoms and atmospheric H2. The produced OH is very energetic,
capable of escaping into space, and expected to contribute a small fraction
of rotationally and vibrationally excited hydroxyl molecules to the Martian
hot corona. These OH molecules are expected to have lifetimes up to several
hours providing they are not destroyed in collisions. We estimate that the
process contributes up to about 0.1% of the total rate of escape of the Martian
atmosphere to space. Our model is based on state-to-state energy dependent
reactive cross sections for O+H2 reaction and a 1D model of transport in
the Martian atmosphere. The hot O atoms produced by photo-dissociative
13
recombination of O+
2 and collisions of thermal oxygen atoms with H and He
ENAs were considered. The ENAs are found to contribute less than 1% to
the total formation rates for the present-day solar flux.
This non-thermal process could be identified by its characteristic emission
profile from high rotational and vibrational states of OH (vibrational states
up to v(cid:48) = 6 could be excited). Recently, Meinel emission bands up to
(3 − 2) and (3 − 1) were detected in limb observations of Mars' atmosphere
with MRO CRISP instrument and taken as a signature of presence of OH
in the Martian atmosphere (Clancy et al., 2013). The emissions are likely to
originate from the altitudes between 45 and 55 km, where the concentration
of the OH is the highest, and are described well by the existing models
that include relevant chemical processes in middle atmosphere. While low
densities of OH in the upper atmosphere present a significant difficulty for
observations, identification of emissions from high Meinel OH bands, with the
intensities following the predicted altitude profile, would confirm the presence
of this non-thermal process. Moreover, the described non-thermal mechanism
and detailed information about populated excited states may be helpful in
interpreting high-resolution spectra of neutrals in the upper atmosphere of
Mars, leading to more accurate estimates of total escape rates of H2 and O
from Mars.
We expect that the developed model can be adapted to other planetary
atmospheres where the described high-temperature reactions can take place,
including comets.
5. Acknowledgments
M.G.'s research was supported by an appointment to the NASA Postdoc-
toral Program at the NASA Ames Research Center, administered by Univer-
sities Space Research Association under contract with NASA.
References
Acuna, M., Connerney, J., Lin, R., Mitchell, D., Carlson, C., McFadden, J.,
Anderson, K., R`eme, H., Mazelle, C., Vignes, D., et al., 1999. Global dis-
tribution of crustal magnetization discovered by the Mars Global Surveyor
MAG/ER experiment. Science 284 (5415), 790 -- 793.
14
Balakrishnan, N., Kharchenko, V., Dalgarno, A., Oct. 1998. Slowing of en-
ergetic O(3P) atoms in collisions with N2. J. Geophys. Res. 1032, 23393 --
23398.
Bougher, S. W., Cravens, T. E., Grebowsky, J., Luhmann, J., Dec. 2015. The
Aeronomy of Mars: Characterization by MAVEN of the Upper Atmosphere
Reservoir That Regulates Volatile Escape. Space Sci. Rev. 195, 423 -- 456.
Bovino, S., Zhang, P., Gianturco, F. A., Dalgarno, A., Kharchenko, V., Jan.
2011. Energy transfer in O collisions with He isotopes and Helium escape
from Mars. Geophys. Rev. Lett. 38, L02203.
Brunetti, B., Liuti, G., Pirani, F., Vecchiocattivi, F., Luzzatti, E., Jun. 1981.
Study of the interactions of atomic and molecular oxygen with O2 and N2
by scattering data. J. Chem. Phys. 74, 6734 -- 6741.
Budzien, S. A., Festou, M. C., Feldman, P. D., Jan. 1994. Solar flux variabil-
ity and the lifetimes of cometary H2O and OH. Icarus 107, 164.
Chassefi`ere, E., Leblanc, F., Sep. 2004. Mars atmospheric escape and evolu-
tion; interaction with the solar wind. Planet. Space Sci. 52, 1039 -- 1058.
Clancy, R., Sandor, B. J., Garc´ıa-Munoz, A., Lef`evre, F., Smith, M. D.,
Wolff, M. J., Montmessin, F., Murchie, S. L., Nair, H., Sep. 2013. First
detection of Mars atmospheric hydroxyl: CRISM Near-IR measurement
versus LMD GCM simulation of OH Meinel band emission in the Mars
polar winter atmosphere. Icarus 226, 272 -- 281.
Deighan, J., Chaffin, M. S., Chaufray, J.-Y., Stewart, A. I. F., Schnei-
der, N. M., Jain, S. K., Stiepen, A., Crismani, M., McClintock, W. E.,
Clarke, J. T., Holsclaw, G. M., Montmessin, F., Eparvier, F. G., Thie-
mann, E. M. B., Chamberlin, P. C., Jakosky, B. M., Nov. 2015. MAVEN
IUVS observation of the hot oxygen corona at Mars. Geophys. Res. Lett.
42, 9009 -- 9014.
Dong, C., Bougher, S. W., Ma, Y., Toth, G., Lee, Y., Nagy, A. F., Tenishev,
V., Pawlowski, D. J., Combi, M. R., Najib, D., Sep. 2015. Solar wind
interaction with the Martian upper atmosphere: Crustal field orientation,
solar cycle, and seasonal variations. J. Geophys. Res. 120, 7857 -- 7872.
15
Fox, J. L., Sep. 1993. On the escape of oxygen and hydrogen from Mars.
Geophys. Res. Lett. 20, 1747 -- 1750.
Fox, J. L., Ha´c, A. B., Dec. 2009. Photochemical escape of oxygen from Mars:
A comparison of the exobase approximation to a Monte Carlo method.
Icarus 204, 527 -- 544.
Gacesa, M., Kharchenko, V., Oct. 2014. Quantum reactive scattering of
O(3P)+H2 at collision energies up to 4.4 eV. J. Chem. Phys. 141 (16),
164324.
Gacesa, M., Zhang, P., Kharchenko, V., May 2012. Non-thermal escape of
molecular hydrogen from Mars. Geophys. Res. Lett. 39, L10203.
Guberman, S. L., Jan. 1988. The production of O(1D) from dissociative re-
combination of O+
2 . Planet. Space Sci. 36, 47 -- 53.
Ip, W.-H., Oct. 1988. On a hot oxygen corona of Mars. Icarus 76, 135 -- 145.
Jakosky, B. M., Lin, R. P., Grebowsky, J. M., Luhmann, J. G., Mitchell,
D. F., Beutelschies, G., Priser, T., Acuna, M., Andersson, L., Baird,
D., Baker, D., Bartlett, R., Benna, M., Bougher, S., Brain, D., Carson,
D., Cauffman, S., Chamberlin, P., Chaufray, J.-Y., Cheatom, O., Clarke,
J., Connerney, J., Cravens, T., Curtis, D., Delory, G., Demcak, S., De-
Wolfe, A., Eparvier, F., Ergun, R., Eriksson, A., Espley, J., Fang, X.,
Folta, D., Fox, J., Gomez-Rosa, C., Habenicht, S., Halekas, J., Holsclaw,
G., Houghton, M., Howard, R., Jarosz, M., Jedrich, N., Johnson, M.,
Kasprzak, W., Kelley, M., King, T., Lankton, M., Larson, D., Leblanc, F.,
Lefevre, F., Lillis, R., Mahaffy, P., Mazelle, C., McClintock, W., McFad-
den, J., Mitchell, D. L., Montmessin, F., Morrissey, J., Peterson, W., Pos-
sel, W., Sauvaud, J.-A., Schneider, N., Sidney, W., Sparacino, S., Stewart,
A. I. F., Tolson, R., Toublanc, D., Waters, C., Woods, T., Yelle, R., Zurek,
R., Dec. 2015. The Mars Atmosphere and Volatile Evolution ( MAVEN)
Mission. Space Sci. Rev. 195, 3 -- 48.
Johnson, R. E., 1982. Introduction to atomic and molecular collisions.
Plenum Press, 1982, New York.
Johnson, R. E., Combi, M. R., Fox, J. L., Ip, W.-H., Leblanc, F., McGrath,
M. A., Shematovich, V. I., Strobel, D. F., Waite, J. H., Aug. 2008. Exo-
spheres and Atmospheric Escape. Space Sci. Rev. 139, 355 -- 397.
16
Kharchenko, V., Dalgarno, A., Fox, J. L., Dec. 2005. Thermospheric distri-
bution of fast O(1D) atoms. J. Geophys. Res. 110, A12305.
Kokoouline, V., Dulieu, O., Kosloff, R., Masnou-Seeuws, F., May 1999.
Mapped Fourier methods for long-range molecules: Application to per-
+) photoassociation spectrum. J. Chem. Phys.
turbations in the Rb2(0u
110, 9865 -- 9876.
Krasnopolsky, V. A., Jun. 2010. Solar activity variations of thermospheric
temperatures on Mars and a problem of CO in the lower atmosphere. Icarus
207, 638 -- 647.
Krest'yanikova, M. A., Shematovich, V. I., Jan. 2005. Stochastic models of
hot planetary and satellite coronas: a photochemical source of hot Oxygen
in the upper atmosphere of Mars. Sol. Syst. Res. 39, 22 -- 32.
Lammer, H., Chassefi`ere, E., Karatekin, O., Morschhauser, A., Niles, P. B.,
Mousis, O., Odert, P., Mostl, U. V., Breuer, D., Dehant, V., Grott, M.,
Groller, H., Hauber, E., Pham, L. B. S., Jan. 2013. Outgassing History
and Escape of the Martian Atmosphere and Water Inventory. Space Sci.
Rev. 174, 113 -- 154.
Lee, Y., Combi, M. R., Tenishev, V., Bougher, S. W., Deighan, J., Schneider,
N. M., McClintock, W. E., Jakosky, B. M., Nov. 2015a. A comparison of
3-D model predictions of Mars' oxygen corona with early MAVEN IUVS
observations. Geophys. Res. Lett. 42, 9015 -- 9022.
Lee, Y., Combi, M. R., Tenishev, V., Bougher, S. W., Lillis, R. J., Nov.
2015b. Hot oxygen corona at Mars and the photochemical escape of oxygen:
Improved description of the thermosphere, ionosphere, and exosphere. J.
Geophys. Res. 120, 1880 -- 1892.
Lewkow, N. R., Kharchenko, V., Aug. 2014. Precipitation of Energetic Neu-
tral Atoms and Induced Non-thermal Escape Fluxes from the Martian
Atmosphere. Astrophys. J. 790, 98.
Lillis, R. J., Brain, D. A., Bougher, S. W., Leblanc, F., Luhmann, J. G.,
Jakosky, B. M., Modolo, R., Fox, J., Deighan, J., Fang, X., Wang, Y. C.,
Lee, Y., Dong, C., Ma, Y., Cravens, T., Andersson, L., Curry, S. M.,
Schneider, N., Combi, M., Stewart, I., Clarke, J., Grebowsky, J., Mitchell,
17
D. L., Yelle, R., Nagy, A. F., Baker, D., Lin, R. P., Dec. 2015. Charac-
terizing Atmospheric Escape from Mars Today and Through Time, with
MAVEN. Space Sci. Rev. 195, 357 -- 422.
McElroy, M. B., Donahue, T. M., Sep. 1972. Stability of the Martian Atmo-
sphere. Science 177, 986 -- 988.
Meinel, I. A. B., May 1950. OH Emission Bands in the Spectrum of the Night
Sky. Astrophys. J. 111, 555.
Nagy, A. F., Winterhalter, D., Sauer, K., Cravens, T. E., Brecht, S., Mazelle,
C., Crider, D., Kallio, E., Zakharov, A., Dubinin, E., Verigin, M., Kotova,
G., Axford, W. I., Bertucci, C., Trotignon, J. G., Mar. 2004. The plasma
Environment of Mars. Space Sci. Rev. 111, 33 -- 114.
Parker, G. A., Pack, R. T., Feb. 1978. Rotationally and vibrationally inelastic
scattering in the rotational IOS approximation. Ultrasimple calculation of
total (differential, integral, and transport) cross sections for nonspherical
molecules. J. Chem. Phys. 68, 1585 -- 1601.
Rahmati, A., Larson, D. E., Cravens, T. E., Lillis, R. J., Dunn, P. A.,
Halekas, J. S., Connerney, J. E., Eparvier, F. G., Thiemann, E. M. B.,
Jakosky, B. M., Nov. 2015. MAVEN insights into oxygen pickup ions at
Mars. Geophys. Res. Lett. 42, 8870 -- 8876.
van Dishoeck, E. F., Dalgarno, A., Sep. 1984. The dissociation of OH and
OD in comets by solar radiation. Icarus 59, 305 -- 313.
Wang, X.-D., Barabash, S., Futaana, Y., Grigoriev, A., Wurz, P., Dec. 2013.
Directionality and variability of energetic neutral hydrogen fluxes observed
by Mars Express. J. Geophys. Res. 118, 7635 -- 7642.
Zhang, P., Kharchenko, V., Jamieson, M. J., Dalgarno, A., Jul. 2009. Energy
relaxation in collisions of hydrogen and deuterium with oxygen atoms. J.
Geophys. Res. 114, A07101.
18
|
1809.01629 | 1 | 1809 | 2018-09-05T17:13:33 | Magma oceans as a critical stage in the tectonic development of rocky planets | [
"astro-ph.EP"
] | Magma oceans are a common result of the high degree of heating that occurs during planet formation. It is thought that almost all of the large rocky bodies in the Solar System went through at least one magma ocean phase. In this paper, we review some of the ways in which magma ocean models for the Earth, Moon, and Mars match present day observations of mantle reservoirs, internal structure, and primordial crusts, and then we present new calculations for the oxidation state of the mantle produced during the magma ocean phase. The crystallization of magma oceans likely leads to a massive mantle overturn that may set up a stably stratified mantle. This may lead to significant delays or total prevention of plate tectonics on some planets. We review recent models that may help partly alleviate the mantle stability issue and lead to earlier onset of plate tectonics. | astro-ph.EP | astro-ph | Phil. Trans. R. Soc. A.
doi:10.1098/not yet assigned
Magma oceans as a critical stage in the tectonic
development of rocky planets
Laura Schaefer1,2,* and Linda T. Elkins-Tanton1
1School of Earth and Space Exploration, Arizona State University, Tempe, AZ 85287, USA
2ORCID: 0000-0003-2915-5025
Keywords: magma oceans, planet formation, differentiation, accretion
Summary
Magma oceans are a common result of the high degree of heating that occurs during planet formation. It is
thought that almost all of the large rocky bodies in the Solar System went through at least one magma ocean
phase. In this paper, we review some of the ways in which magma ocean models for the Earth, Moon, and
Mars match present day observations of mantle reservoirs, internal structure, and primordial crusts, and then
we present new calculations for the oxidation state of the mantle produced during the magma ocean phase.
The crystallization of magma oceans likely leads to a massive mantle overturn that may set up a stably
stratified mantle. This may lead to significant delays or total prevention of plate tectonics on some planets. We
review recent models that may help partly alleviate the mantle stability issue and lead to earlier onset of plate
tectonics.
Magma oceans are common
Magma oceans appear to be a common outcome of formation processes for large rocky bodies [1,2]. Small
planetesimals and embryos can form on very short timescales in the protoplanetary disk, so that they
incorporate significant amounts of short-lived radionuclides such as 26Al and 60Fe that can produce enough
heat to at least partially melt many of these objects [3,4]. Models suggest that the asteroid 4 Vesta, which is the
source of the howardite-eucrite-diogenite (HED) meteorites, went through a magma ocean stage due to such
short-lived heating [5-7]. Many iron meteorites, which are likely the remnants of differentiated planetesimals,
have ages of only ~1 Myr after Solar System formation, indicating very early, short-lived magma oceans [8,9].
*Author for correspondence ([email protected]).
2
Larger objects like the Earth take long enough to assemble that short-lived radionuclides cannot provide
sufficient heating to melt them. For these larger, later-assembling bodies, substantial heating can come from
accretionary impacts [e.g. 10,11] and gravitational segregation of metallic iron into the centre of the planet.
Giant impacts also appear to be common in N-body simulations of rocky planet formation [12], and are likely
to cause wide-spread melting that can lead to differentiation in both silicates and metals [13].
In this paper we will discuss several ways in which forward magma ocean models do a good job of
matching real-world observations of the Earth, Moon, and Mars. Magma ocean models can produce large,
low-shear velocity provinces (LLSVPs) at the base of Earth's mantle through cumulate overturn and predict
that these structures are stable over billions of years. Lunar magma ocean models can produce the internal
differentiation of the Moon, including the anorthositic crust, the source regions for the picritic glasses and
mare basalts, and the KREEP component, rich in incompatible elements. Models of the magma ocean on Mars
can produce early crust, some of which may be preserved to the present day, and which could be hydrated by
the earliest atmosphere to produce primordial clays. We also introduce new calculations that show that
magma ocean models can produce the present-day mantle oxidation state of the Earth through metal-silicate
equilibration within the magma ocean, with no additional mechanisms required. We will then discuss the
mantle state at the end of the magma ocean period and discuss whether it can hinder or allow early onset of
plate tectonics.
Magma ocean models match observations
Magma ocean models have been applied to all of the terrestrial planets, as well as the Earth's Moon. These
models match observable properties of the present-day planets. The case for magma oceans on the Moon and
Mars are easier to make because of the lack of active tectonic processing over most of the Solar System history.
For the Earth, models can match deep mantle structures and the chemical properties of the mantle, including
the abundances of trace elements and the present-day oxidation state. We review these model predictions
below and present new calculations of the oxidation state of the magma ocean during core formation on the
Earth.
Differentiation of the Lunar Interior
Magma ocean models were first fully developed to explain observations and measurements of Apollo
returned samples from the Moon indicating that there was abundant plagioclase feldspar in the lunar crust
[14-16]. The feldspar found in the lunar crust is nearly pure (> 98%) anorthosite, containing trace amounts of
Phil. Trans. R. Soc. A.
3
mafic phases [15,17,18]. Lunar meteorites, remote spectroscopy and gravity field studies all support the wide-
spread distribution of anorthosite in the lunar crust [19-21]. This widespread occurrence has led to the
hypothesis that the Moon was at least partially molten and that anorthite plagioclase buoyantly segregated
from the melt to form the crust [14-16]. This early silicate differentiation event is widely believed to have been
a lunar magma ocean, which would be a natural consequence of the Moon-forming impact. In addition to the
anorthositic crust, the existence of the highly incompatible element enriched KREEP (K, Rare Earth Element
and P -rich) component, as well as the positive Eu anomalies in the highlands anorthosites and
complementary negative Eu anomalies in mare basalts [e.g., 22-24] provide additional evidence in favour of a
global lunar magma ocean. Models of metal-silicate partitioning during lunar core formation are also
consistent with equilibration occurring at the lunar CMB, indicating a fully molten Moon [25].
Numerous authors have modelled the crystallization of the lunar magma ocean [e.g., 13, 26-33]. The
lunar magma ocean would initially be fully molten at the surface and could rapidly cool and solidify by ~80%
by volume in the first 1 ky [33]. The first solids would include Mg-rich olivine followed by orthopyroxene in
the lower lunar mantle. Above 70-80% solidification, plagioclase begins to crystallize, and since it is lower in
density than the magma ocean liquid, it should have floated to the surface to form a thermally conductive lid.
Expulsion of trapped liquid by compaction and other processes would limit the amount of mafic minerals that
could have formed in the crust, in agreement with observations [e.g., 34,35]. The magma ocean melt becomes
more Fe-rich as crystallization proceeds, with a very dense ilmenite-bearing layer forming after about 95%
crystallization. The residual liquids are very enriched in incompatible elements and have compositional
characteristics similar to those found in the KREEP component [30]. Thermal models show that the conductive
anorthositic crust (~40-50 km) would have delayed the solidification of the remaining magma ocean by ∼10
My [33]. However, ages for lunar crustal materials, discussed below, suggest that magma ocean crystallization
was not complete until ~150 -- 200 Myrs. Recent models have shown, however, that tidal heating may extend
the crystallization of this final magma ocean liquid out to 200 Myr [36,37]. Gravitationally driven cumulate
overturn could produce later episodes of volcanism that would explain the ages and compositions of the Mg-
suite of lunar crustal rocks [33].
Magma ocean crystallization experiments [32, 38, 39] reproduce many aspects of the numerical
magma ocean models, including increasing density of the cumulates with crystallization, but some
discrepancies remain. In particular, although liquid evolution trends are broadly similar between experiments
Phil. Trans. R. Soc. A.
4
and models, the late stage liquid composition can differ significantly [Gaffney et al. in prep]. Recent
experiments suggest that a dry lunar magma ocean would produce much thicker anorthositic crust than
observed by the GRAIL mission, but a wet lunar magma ocean containing 270-1650 ppm of water would
produce a thinner crust [38,39]. This water abundance is consistent with recent measurements of water
abundances in lunar anorthosites [40]. Additional models including the effect of water on the crystallization
time will be needed to compare with lunar geochronology measurements discussed below in order to confirm
a wet lunar magma ocean.
The age of lunar crustal anorthosites provides a constraint on the crystallization age of the lunar
magma ocean, but measured ages cover a large span, from 4.47 to 4.29 Gyr [e.g., 41-46]. Kleine et al. [47]
reviewed the isotopic ages for lunar samples available at the time and concluded that the combined
constraints from the Hf-W and 147Sm-143Nd systems indicated that the lunar magma ocean reached ~70%
crystallization, at which point anorthosite flotation should begin, no earlier than 60 Myrs and no later than 150
Myrs. Borg et al. [48] reported a very young age of crystallization of a ferroan anorthosite of 4360 +/- 3 million
years from combined analysis of the 207Pb-206Pb, 147Sm-143Nd and 146Sm-142Nd isotopic systems, making this the
most reliable FAN age, and suggested that the lunar crust was produced by non-magma ocean processes such
as serial magmatism. Boyet et al. [49] measured Sm-Nd systematics in several lunar ferroan anorthosite
samples and found that they did not form a single isochron, which the authors suggest indicates that all but
one of these samples formed through later magmatism or processing after the lunar magma ocean
crystallization. The one sample with the largest deficit in 142Nd gave a crystallization age of 60-125 Myr after
Solar System formation for lunar crust formation. Recent results from Pb-isotopic evolution in lunar basalts
has suggested a large differentiation event at 4376 ± 18 Myrs, which could coincide with the last stages of
lunar magma ocean crystallization [50], consistent with the 142Nd ages for FAN 60025 [48]. Ages determined
for lunar zircons indicate a younger age of 4.51 Gyrs for lunar crustal formation [51], which is significantly
older than ages from ferroan anorthosites. This age, which is closer to Moon-formation ages of ~ 60 Myr from
Hf-W [47], may instead simply represent the time of zircon formation in the initially rapidly cooling lunar
magma ocean [52]. Modelling by Meyer et al. [36] suggests that prolonged cooling of the crust due to tidal
heating could explain the late ages determined for the lunar anorthosites relative to the much shorter expected
cooling times [e.g. 33].
One significant observation that cannot readily be explained by lunar serial magmatism, and therefore
requires a deep magma ocean, is the apparent ubiquity of an orthopyroxene+olivine melting source region for
Phil. Trans. R. Soc. A.
5
the picritic glasses between ~250 and ~500 km depth on the lunar nearside (data summary in [33], Figure 6).
This thick deep layer of uniform mineralogy has to have been created by a deep, fractionating magma ocean:
Serial diapirs will not leave behind a uniform cumulate pile, and the heterogeneities from between the
putative melting regions would be enriched and therefore melt at lower temperatures later, producing melt
products different from the picritic glasses, and not sampled on the Moon.
Further, the lunar crust's thickness is smoothly varying from the nearside to the far side, excepting the
effects of large impacts. To produce a smoothly varying crustal thickness via serial magmatism, the
wavelength of the diapirs must not be more than a few multiples of the crustal thickness, or relaxation of the
rapidly-cooling and high-viscosity anorthosites would not produce a smoothly varying crustal thickness. Such
ubiquity of melting around the Moon must be virtually indistinguishable from a global magma ocean.
Many aspects of the lunar magma ocean remain to be investigated, including the depth of the magma
ocean, partial equilibrium vs fully fractional crystallization [30,33], the cooling times of the crust and the
possible formation of a crustal dichotomy between near and far-side of the Moon [53, 54]. While many details
of the chronology of lunar geology remain to be explained, the lunar magma ocean model is very successful in
producing many of the observed properties of the Moon's internal structure and surface composition.
Martian crust and clay formation
An early Martian magma ocean is supported by measurements of short-lived radioactive isotopes that
indicate early differentiation of Mars and generation of enriched and depleted mantle reservoirs [9, 55-57].
Young core formation ages for Mars (>~4-10 Myrs) [58] are further argument for a magma ocean because
short-lived radioactive isotopes such as 26Al would still have been present to provide additional heat and core
segregation requires at least partial melting of the silicate mantle. However, it remains uncertain how deep the
magma ocean on Mars would have been. Physical models of the accretion and differentiation of Mars have
attempted to determine the degree of melting with variable results. Models using impact heating and cooling
but no other heat source find only partial melting of the near surface and a cold interior [59]. A multiphase
model including radioactive and gravitational heating found that core formation may be a catastrophic, early
event that can lead to additional heating of Mars-sized objects, although full silicate melting is not required
[60]. The blanketing effect of a hybrid proto-atmosphere (both outgassed and nebular components) could also
help to sustain a global magma ocean throughout accretion, depending on the accretion rate of proto-Mars
Phil. Trans. R. Soc. A.
6
[61]. No model including all of the physical processes controlling the heating and cooling of proto-Mars
(impact heating & cooling, radioactive decay, atmospheric blanketing, gravitational segregation, thermal
convection, etc.) exist, and the physical partitioning of energy of those processes remains largely unknown, so
the depth of the martian magma ocean remains uncertain.
The magma ocean stage may have produced silicate reservoirs enriched and depleted in incompatible
elements during fractional crystallization, which remain today due to incomplete mantle mixing. The ages of
these reservoirs as measured with short-lived radioactive isotope systems have been used to infer the duration
of the magma ocean stage and timing of crust formation, although interpretations within the community can
vary. Measurements of Sm-Nd isotope systematics in shergottites by [62] suggest that the magma ocean stage
on Mars lasted for ~100 Myrs, with a depleted reservoir forming by 4.535 Gyr and an enriched reservoir,
plausibly either the crust of late stage magma ocean liquids, forming by 4.457 Myrs. Lead-lead ages of zircons
in the polymict breccia NWA 7034/7355, which formed through melting of primary crust, give maximum
crustal formation times of 4.428 Gyr [63]. Whole rock Sm-Nd ages of the same meteorite give similar ages of
4.42 - 4.46 Gyr [64]. Co-existing variations of ε182W and ε142Nd have been attributed to multiple episodes of
silicate differentiation [9]. These authors find formation times of the depleted reservoir of ~25 Myrs, with
ongoing formation of the enriched reservoir (thought to be the crust), extending to at least 40 Myrs after Solar
System formation. This is ~60 Myrs younger than [62] and ~100 Myrs younger than [63, 64] but is more
consistent with results from martian crust formation models [65]. In comparison, magma ocean thermal
models predict crystallization timescales < 5 Myr for Mars [66, 67]. Additional analyses of Sm-Nd isotope
systematics in shergottites and nahklites found a silicate differentiation time of 4504 Myr (~60 Myrs after Solar
System formation) [68]. Given the discrepancy with magma ocean thermal models [66, 67], these authors
suggest that this age could be explained by a giant impact and heterogeneous mixing of the martian mantle,
rather than a global silicate differentiation event. However, recent high-precision isotopic work on zircons by
[69] shows that the Martian magma ocean was solidified in less than 10 Myr after planetary formation, a result
that indicates that a solid Mars existed within 20 Myr of CAIs. This result supports the rapid solidification of
the Martian magma ocean and indicates that later reservoir disruptions were likely due to giant impacts and
remixing.
Elkins-Tanton et al. [65] modelled early crust formation on Mars at 30-50 Myrs after accretion arising
during cumulate overturn by melting of the hot cumulates rising into the upper mantle. Measurements of
potentially ancient crust in crater central peaks find primarily dunite and pyroxenite-dominated lithologies.
Phil. Trans. R. Soc. A.
7
Compositional variations are suggestive of high degrees of fractional crystallization in a slowly cooling
magma body, potentially generated by the overturn of the mantle following magma ocean solidification [70].
However, [71] find that stagnant lid convection models beginning from a thermal and density state derived
from fractional crystallization cannot reproduce the later long-term thermal evolution of the planet (e.g. 1-2
long-lived mantle plumes). Therefore, more work is still needed in understanding the depth and longevity of
the martian magma ocean, and the post-overturn thermal state.
Clays are relatively common on the surface of Mars [72], suggesting the widespread presence of water
on the surface (or near-surface) throughout the Noachian. Their presence has been used to try to constrain the
climate history of early Mars, with models ranging from 'warm and wet' to 'cold and transiently wet' [73,74].
However, magma ocean models provide another possible source for clay minerals. Magma ocean outgassing
would result in a thick atmosphere, of which a large or primary component, is water vapor [e.g. 66]. Primitive
clays found on the surface of Mars could have formed through rapid reaction of a supercritical steam
atmosphere or hot early ocean during the late magma ocean stage with the earliest magma ocean crust [75].
Although this clay layer would later be buried by new crust [66], this model may account for the presence of a
thick clay layer at large depths within the Martian crust observed in crater central peaks [76] and the
abundance of clay minerals within the regolith. The clay formation model assumes the long-term stability of
the crust against the overturn, which should be further investigated. Early clays would likely have lower D/H
ratios, since they would have formed before significant water escape from the planet, which provides a
testable prediction of the primordial clay model.
Chemical Heterogeneities in the Earth's Lower Mantle
Seismic studies have long noted large scale heterogeneities in the lower mantle from variations in seismic
velocities [77,78]. Both seismically fast and slow regions have been identified. While the seismically fast
regions seem to be associated with regions of subduction, the seismically slow regions --called large low shear
velocity provinces, or LLSVPs -- are found in regions associated with hot spot volcanism [79,80]. Seismic
tomography suggests that the LLSVPs are associated with chemical heterogeneity in the lower mantle, and
that these regions have higher densities than the surrounding lower mantle [81,82]. Within the LLSVPs are
small (5-40 km thick) regions of even lower seismic velocity (ultralow-velocity zones, ULVZs), that have been
associated with partial melting of silicates directly above the core or with significant compositional differences
[83 -86]. Geodynamics simulations suggest that the LLSVPs could have been stable throughout Earth's history
Phil. Trans. R. Soc. A.
8
if they are higher in density and/or viscosity than the surrounding lower mantle [87-89], suggesting a potential
primordial origin.
Chemical heterogeneities within the mantle suggest the existence of a small enriched mantle reservoir
in the deep mantle [43, 90]. This suggestion is based on the Earth's 142Nd/144Nd ratio, which is higher than
measured in chondrites and suggests that an early silicate differentiation event occurred in the first 160 Myrs
[91]. Other recent measurements [92] suggest that the Earth has the same 142Nd/144Nd ratio as enstatite
chondrites, obviating the need for an enriched reservoir. Other arguments in favour of a primordial chemical
reservoir in the mantle come from noble gases. Mid-ocean ridge basalts (MORBs) and ocean island basalts
(OIBs) appear to tap different chemical reservoirs, with the MORB source being depleted in primordial noble
gas components and the OIB source being enriched and potentially undegassed [93-97].
Fractional crystallization of a magma ocean produces an unstable density stratification due to the
enrichment of Fe in late stage melts [1, 66, 98-100]. These models predict mantle overturn following the
crystallization of the magma ocean, with the final iron-rich residues sinking to the core mantle boundary of
the planet. Since they are denser than the overlying mantle, these overturned cumulates might be stable at the
CMB on long time scales [e.g. 101], so models suggest that these could be responsible for the modern day
LLSVPs [102].
Alternative primordial origins for the LLSVPs have been suggested, including a primary basal magma
ocean [103], partial melting of the Archaean mantle and sinking of dense, Fe-rich melts [104], or subduction of
early crust enriched with late delivered chondritic material [94105 A primary basal magma ocean is produced
if the liquidus and the magma ocean adiabat first intersect in the middle of the mantle, with crystallization
proceeding towards the surface rapidly, and more slowly towards the core. The basalt melt would represent
an un-depleted primordial chemical reservoir, which could generate denser crystallization products as Fe
becomes enriched in the melt [103]. Zhang et al. [106] alter the basal magma ocean model to suggest that the
higher density of the LLSVPs is due to the trapping of Fe-Ni-S metallic liquids within the basal magma ocean
cumulates. However, recent analysis of mantle melting curve [107 ] suggests the mantle liquidus should
increase smoothly from surface to core-mantle boundary (CMB). If this is verified by further experiments,
magma ocean crystallization would initiate at the CMB, and a primary basal magma ocean could likely be
ruled out.
Phil. Trans. R. Soc. A.
9
Figure 1
Oxidation state of Earth's mantle through geologic time from redox proxy measurements. See text for source
details. The two data points at ~ 0 Myrs are values for present day MORBs. We plot historical oxygen fugacity
relative to the fayalite-magnetite-quartz buffer (FMQ). At 1200C and 1 GPa, NNO = FMQ + 0.7, IW = FMQ --
3.69.
Mantle oxidation state
The oxidation state of the upper mantle of the present-day Earth is near the FMQ buffer [108]. Various
estimates are of the mantle redox state through time are shown in Figure 1. Measurements of ancient redox
proxies, Cr and V whole-rock abundances in ancient volcanics and the composition of Cr-rich spinels in
ancient volcanics indicate that the oxygen fugacity of the mantle has been at or above the present-day value
(±0.5 log fO2) since ~3900 Ma [109]. Canil [110] used partitioning of V between komatiitic liquids and olivine in
6 well-characterized komatiite flows to estimate Archaean fO2 and found values from ΔNNO -2 to +1. Li & Lee
[111] looked at V/Sc ratios in basalts and find that Archaean basalts have V/Sc ratios only slightly smaller than
Phil. Trans. R. Soc. A.
10
that of modern MORBS, indicating only a minor decrease of ~0.3 log fO2 units relative to FMQ. Recent
measurements of 3.5 -- 2.4 Ga komatiites used V partitioning behaviour to constrain the fO2 of the mantle, and
found a range of -0.11 to +0.43 FMQ, consistent with a relatively constant mantle fO2 through time [112].
Aulbach & Stagno [113] used V/Sc ratios in metamorphosed mid-ocean ridge basalts (MORBs) and picrites
from 3.0 -- 0.55 Ga as redox proxies and found a slightly larger decrease in mantle fO2, down to -1.5 FMQ,
during the Archean. They note that the increase of the mantle fO2 to the present-day value appears coincident
with the atmospheric Great Oxidation Event (GOE). Analysis of Ce abundances in detrital Hadean-aged Jack
Hills zircons suggest oxygen fugacities consistent with the present day back to 4.3 Ga [114]. Taken together,
the present measurements are consistent with a constant oxygen fugacity within the mantle over the entire
history of the Earth.
However, models of core formation, which occurs within the magma ocean of the forming Earth,
require much more reducing conditions in order to match trace element partitioning behaviour between the
mantle and the core [e.g. 115-117]. Therefore, oxidation of the mantle to the present-day state had to occur
contemporaneously with core formation or shortly thereafter and operate on a relatively short timescale.
Several models have been proposed to increase the oxidation state of the mantle through the post-production
of Fe3+ in the mantle. These include gradual H escape from accreted water [118], and the production of Fe3+
through disproportionation of Fe2+ in the lower mantle by crystallization of bridgmanite [119,120]. Trace
element partitioning models [e.g. 117, 121] match the present-day mantle FeO abundance by allowing the
composition of accreting material to become more oxidized through time (IW-5 to IW-2), but these models fail
to predict Fe3+/ΣFe, which governs present-day oxidation state. These models typically end with a mantle at
IW-2, about 5 log fO2 units below the present day.
We propose an alternative mechanism: that disproportionation of Fe2+ occurs in the silicate melt phase
in the magma ocean during equilibration with the core forming metal delivered by accreting impactors, rather
than during crystallization. This allows immediate separation of the metallic liquid, avoiding the remixing
problem of [119,120]. The equilibrium amongst the different Fe valence states will be given by the reaction
3FeO(melt) -> 2FeO1.5(melt) + Fe (liquid metal). The metal will sink through the magma ocean to join the core,
leaving the oxidized mantle material behind, without requiring complicated scenarios of melting and
recrystallizing of bridgmanite that would be difficult to quantify. To reproduce the present day upper mantle,
metal-silicate equilibration in the melt at high pressures must produce Fe3+/ΣFe ~ 0.01 -- 0.05 [108]. As in [122],
we assume that equilibrium will be controlled at the level of metal-silicate equilibration and the magma ocean
Phil. Trans. R. Soc. A.
11
Figure 2
Fe3+/ΣFe in the magma ocean produced by equilibration of liquid silicate and core-forming metal as a function
of pressure along the mantle liquidus of [123]. See Table 1 for thermodynamic data sources. The grey shaded
region highlights the Fe3+/ΣFe values in the Earth's present day upper mantle [108].
composition would be rapidly homogenized by convection. In Figure 2, we show the amount of Fe3+ produced
through this reaction as a function of temperature along the mantle liquidus [123], using different data sets for
partial molar volumes of the molten iron oxides [124, 125]. The low-pressure model data [124, 126] was
measured at 1 bar, whereas for the higher-pressure model, we use data from new measurements of the
Fe3+/ΣFe ratio in high pressure melts [125] of up to 7 GPa to refit the partial molar volume of the Fe3+ oxide. We
use the equation of state of liquid Fe metal from [127] for all calculations and include partitioning of oxygen
and silicon into the metal phase using the partitioning model of [128]. The thermodynamic data used for these
calculations are given in Table 1. We calculate the Fe3+/ΣFe ratio as a function of pressure up to 50 GPa, which
is close to average conditions of core formation determined for the Earth from trace element partitioning
studies [117,121].
The updated equation of state for the liquid iron oxide from [125] predicts much larger Fe3+
abundances are produced at high pressures than the 1 bar data of [124] do. This is consistent with new,
Phil. Trans. R. Soc. A.
12
Table 1.
Fit parameters used to calculate the Fe3+/ΣFe ratio as a function of pressure (see Figure 2). We refit the data of
[125] to a Murnaghan equation of state for FeO1.5 (liq), while taking values from [124] and [126] for FeO(liq). For
both models, we take the standard Gibbs energy of formation of the liquid oxides from [148] and calculate
activity coefficients for the liquid oxides from [149]. For both models, we take a value of κ' = 4 for both oxides.
Low P model
Reference
High P model
Reference
𝑉 FeO (cm3/mole)
𝑉 FeO1.5 (cm3/mole)
𝜅𝐹𝑒𝑂(GPa)
𝜅𝐹𝑒𝑂1.5 (GPa)
13.65
21.065
30.33
16.6
[124]
[124]
[126]
[126]
13.65
18.9
30.33
10.8
[124]
Fit to [125]
[126]
Fit to [125]
unpublished measurements of Fe3+/ΣFe in silicate melts from [129], which go up to pressures of ~25 GPa. The
stabilization of Fe3+ in the melt phase at high pressure was predicted by [122], partially on the basis of the
stability of Fe3+ in the crystalline phase at high pressures. As Fig. 2 shows, the equilibration of the different Fe
valences at high pressure during the separation of core forming metal can easily produce the observed present
day Fe3+ abundances in the Earth's upper mantle, shown by the grey shaded region. Therefore, no additional
oxidation mechanisms are required to produce the present-day oxidation state, which implies that there has
been limited oxidation of the Earth's upper mantle over geologic time. The calculations with the new data also
suggest that the mantles of larger planets should in general be more oxidized than smaller planets once the
metallic phase has separated, given that more Fe3+ is produced at higher pressures. This is corroborated by the
low initial oxygen fugacity of the depleted martian mantle (~IW) [130], in comparison to the Earth's much
higher oxygen fugacity. This has significant implications for the compositions of outgassed atmospheres we
should expect on rocky exoplanets of different sizes.
Magma oceans set the initial conditions for planetary tectonics
Density stratification and magma ocean overturn
Models of fractional crystallization of magma oceans result in unstable density structures due to the
enrichment of late-stage melts in iron. These crystallization models predict whole-mantle overturn at the end
of magma ocean crystallization, which results in a stably stratified density structure in the earliest solid
mantle. This stable stratification is predicted to inhibit the onset of whole mantle convection, which may result
Phil. Trans. R. Soc. A.
13
in an early stagnant lid tectonic regime on rocky planets [131, 132]. Recent models [133, 134] investigate the
possibility of solid state convection beginning within the cumulate pile before the magma ocean has fully
crystallized. Ballmer et al. [133] investigate incremental overturns within the solidifying cumulate pile using
2D numerical convection simulations and find that the majority of the mantle becomes well mixed. However,
the final Fe-rich cumulate melt layer sinks to the CMB during the final overturn and remains stable as a highly
dense layer for geologic time, which may be related to present day LLSVPs. Maurice et al. [134] similarly find
that early onset of solid state convection while the magma ocean is still crystallizing will more efficiently
homogenize the mantle than late stage overturn. However, their model fails to produce a LLSVP-like zone at
the base of the mantle, and they instead prefer the basal magma ocean model of [103]. Tikoo & Elkins-Tanton
[135] show that water within the late-stage melt may be expelled during mantle overturn, which would cause
a large influx of water into the upper mantle. The large abundance of water in the early upper mantle would
have allowed partial melting and lowered the viscosity, possibly allowing the early initiation of whole mantle
convection. These models show that the mantle structure at the end of magma ocean phases remains
uncertain.
Tectonics after the magma ocean
Recent work in scaling theory, steady-state simulations, and numerical evolutionary models show that the
initial conditions of a solid-state mantle convection simulation can strongly influence the tectonic style of the
simulated planet [136-138]. For instance, [136] demonstrate that tectonic models show a history dependence,
i.e. the tectonic mode of a system is determined by its specific geologic and climatic history. Therefore, the
density and thermal structure at the end of the magma ocean stage and following cumulate overturn sets the
stage for the future tectonic evolution of the planet. As discussed in the previous section, a number of
parameters remain poorly constrained by the end stage of magma ocean models, which influences the later
tectonic models. The uncertain parameters include the mantle potential temperature and heat flux, the degree
of mantle hydration, the thickness of the lithosphere, and the planetary climate.
After discovering remnant magnetism on Mars, models predicting early plate tectonics were popular
to explain rapid core cooling needed for magnetic field generation [139]. However, evidence of significant
mantle heterogeneities on Mars suggest that significant convective mixing has not occurred. There is also little
evidence of crustal recycling, indicating that Mars likely entered a stagnant or sluggish-lid phase following the
magma ocean [71, 132], with heat fluxes out of the mantle dominated by a few long-lived mantle plumes.
Phil. Trans. R. Soc. A.
14
It remains uncertain if plate tectonics on the Earth can initiate immediately after the magma ocean
stage, or if the magma ocean is more likely to be followed by a stagnant-lid mode [140] or alternative tectonic
regime (e.g. squishy lid [141], heat pipe model [142], etc.). Foley et al. [143] investigate onset of subduction
during the Hadean using a model including grain-damage and find that a period of stagnant lid convection
follows the initial mantle overturn, and that sluggish subduction can initiate relatively rapidly, due to damage
build-up in the cold, dense lid material. Recent models have highlighted the potential role of triggers for
initiating subduction, including external impacts that punch through the lithosphere and mantle plumes [144,
145]. Numerical models of core-formation suggest that metal diapirs may set up conduits to permit early onset
of mantle plumes, either purely thermal or thermochemically driven [146, 147], which is particularly effective
if the magma ocean is restricted to the upper mantle. These early mantle plumes could then play a role in
triggering onset of early subduction.
This discussion highlights the uncertainty in the tectonic state of planets following the magma ocean
stage and the need for additional modelling. Magma ocean models must in future be more coupled with
petrologic data from short-lived isotopic systems to help us better understand the early evolution of the
terrestrial planets and the requirements for the onset of plate tectonics. Plate tectonics is often thought to be
essential for the sustainability of life, and therefore such work has implications for the search for life on
exoplanets as well as in our own Solar System.
Additional Information
Data Accessibility
The dataset necessary to reproduce Figure 2 have been uploaded as part of the Supplementary Material.
Sources of data shown in Figure 1 are given in the text.
Authors' Contributions
LS and LTET conceived of and designed the study, and LS drafted the manuscript. All authors read and
approved the manuscript.
Competing Interests
The authors declare that they have no competing interests.
Funding Statement
The authors acknowledge funding from Arizona State University's Interplanetary Initiative, the NASA Psyche
mission and the NSF CSEDI program.
Phil. Trans. R. Soc. A.
15
References
[1] Solomatov, V 2000. Fluid dynamics of a terrestrial magma ocean. In:Origin of the Earth and Moon. Ed.
Canup, RM, Righter, K. Univ. AZ Press, 323-338.
[2] Elkins-Tanton, LT 2012. Magma oceans in the Inner Solar System. Ann. Rev. Earth Planet. Sci. 40, 113-39.
[3] Miyamoto M, Fujii N, Takeda H. 1981. Ordinary chondrite parent body: an internal heating model. Proc.
Lunar Planet. Sci. Conf. 12, 1145 -- 52.
[4] Hevey, PJ, Sanders, IS 2006. A model for planetesimal meltdown by 26Al and its implications for meteorite
parent bodies. Met. Planet. Sci. 41, 95-106.
[5] Papanastassiou DA, Wasserburg GJ. 1969. Initial strontium isotopic abundances and the resolution of small
time differences in the formation of planetary objects. Earth Planet. Sci. Lett. 5, 361 -- 76
[6] Neumann, W, Breuer, D, Spohn, T 2014. Differentiation of Vesta: implications for a shallow magma ocean.
Earth Planet. Sci. Lett. 395, 267-280.
[7] Mandler, BE, Elkins-Tanton, LT 2013. The origin of eucrites, diogenites, and olivine diogenites: magma
ocean crystallization and shallow magma chamber processes on Vesta. Met. Planet. Sci. 48, 2333-2349.
[8] Scherstén A, Elliott T, Hawkesworth C, Russell S, Masarik J. 2006. Hf-W evidence for rapid differentiation
of iron meteorite parent bodies. Earth Planet. Sci. Lett. 241, 530 -- 42
[9] Kruijer, TS, Kleine, T., Borg, LE, Brennecka, GA, Irving, AJ, Bischoff, A, Agee, CB. 2017. The early
differentiation of Mars inferred from Hf-W chronometery. Earth Planet. Sci. Lett. 474, 345-354.
(http://dx.doi.org/10.1016/j.epsl.2017.06.047)
[10] Matsui, T, Abe, Y 1986. Formation of a 'magma ocean' on the terrestrial planets due to the blanketing
effect of an impact-induced atmosphere. Earth, Moon, Planets 34, 223-230.
[11] Abe, Y 2011. Protoatmospheres and surface environment of protoplanets. Earth Moon Planets 108, 9-14.
[12] Quintana, EV, Barclay, T, Borucki, WJ, Rowe, JF, Chambers, JE 2016. The frequency of giant impacts on
Earth-like worlds. Astrophys. J. 821, 126.
[13] Tonks, WB, Melosh, HJ 1993. Magma ocean formation due to giant impacts. J. Geophys. Res. 98, 5319-5333.
[14] Smith, J.V., Anderson, A.T., Newton, R.C., Olsen, E.J., Wyllie, P.J., Crewe, A.V., Isaacson, M.S., Johnson,
D., 1970. Petrologic history of the moon inferred from petrography, mineralogy and petrogenesis of Apollo 11
rocks. Proc. Apollo 11 Lunar Sci. Conf. 1, 897 -- 925.
[15] Warren, PH. 1985. The magma ocean concept and lunar evolution. Ann. Rev. Earth Planet. Sci. 13, 201-240.
Phil. Trans. R. Soc. A.
16
[16] Wood, J.A., Dickey, J.S.J., Marvin, U.B., Powell, B.N., 1970. Lunar anorthosites and a geophysical model of
the moon. Proc. Apollo 11 Lunar Sci. Conf., 965 -- 988.
[17] Donaldson-Hanna, KL, Cheek, LC, Pieters, CM, Mustard, JF, Greenhagen, BT, Thomas, IR, Bowles, NE.
2014. Global assessment of pure crystalline plagioclase across the Moon and implications for the evolution of
the primary crust. J. Geophys. Res. Planets, 119, 1516-1545.
[18] Ohtake, M and 37 others. 2009. The global distribution of pure anorthosite on the Moon. Nature, 461, 236-
240.
[19] Korotev, RL, Jolliff, BL, Zeigler, RA, Gillis, JJ, Haskin, LA. 2003. Feldspathic lunar meteorites and their
implications for compositional remote sensing of the lunar surface and the composition of the lunar crust.
Geochim. Cosmochim. Acta 67, 4895-4923.
[20] Hawke, BR, Peterson CA, Blewett, DT, Bussey, DBJ, Lucey, PG, Taylor, GJ, Spudis, PD 2003. Distribution
and modes of occurrence of lunar anorthosite. J. Geophys. Res Planets 108, 5050. (10.1029/2002JE0018900)
[21] Besserer, J, Nimmo, F, Wieczorek, MA, Weber, RC, Kiefer, WS, McGovern, PJ, Andrews-Hanna, JC, Smith,
DE, Zuber, MT. 2014. GRAIL gravity constraints on the vertical and lateral density structure of the lunar crust.
Geophys. Res. Lett. 41, 5771-5777.
[22] Haskin, LA, Lindstrom, MM, Salpas, PA, Lindstrom, D, 1981. On compositional variations among lunar
anorthosites. Proc. Lunar Planet. Sci. Conf. 12, 41 -- 66.
[23] Morse, SA, 1982. Adcumulus growth of anorthosite at the base of the lunar crust. Proc. Lunar Planet. Sci.
Conf. 13, A10 -- A18.
[24] Ryder, G, 1982. Lunar anorthosite 60025, the pertrogenesis of lunar anorthosites, and the bulk
composition of the Moon. Geochem. Cosmochim. Acta 46, 1591 -- 1601.
[25] Rai, N, van Westrenen, W. 2014. Lunar core formation: new constraints from metal-silicate partitioning of
siderophile elements. Earth Planet. Sci. Lett. 388, 343-352.
[26] Longhi, J, 1980. A model of early lunar differentiation. Proc. Lunar Planet. Sci. Conf., 289 -- 315.
[27] Longhi, J, 2003. A new view of lunar ferroan anorthosites: postmagma ocean petrogenesis. J. Geophys. Res.
108, 2-1 -- 2-15.
[28] Longhi, J, 2006. Petrogenesis of picritic mare magmas: Constraints on the extent of early lunar
differentiation. Geochim. Cosmochim. Acta 70, 5919 -- 5934.
[29] Taylor, SR, Jakes, P, 1974. The geochemical evolution of the Moon. Proc. Lunar Sci. Conf. 5, 1287 -- 1305.
Phil. Trans. R. Soc. A.
17
[30] Snyder, GA, Taylor, LA, Neal, CR, 1992. A chemical model for generating the sources of mare basalts:
combined equilibrium and fractional crystallization of the lunar magmasphere. Geochem. Cosmochim. Acta 56,
3809 -- 3823.
[31] Longhi J, Durand SR, Walker D. 2010. The pattern of Ni and Co abundances in lunar olivines. Geochim.
Cosmochim. Acta 74, 784 -- 798.
[32] Elardo, SM, Draper, DS, Shearer, CK, 2011. Lunar Magma Ocean crystallization revisited: Bulk
composition, early cumulate mineralogy, and the source regions of the highlands Mg-suite. Geochim.
Cosmochim. Acta 75, 3024 -- 3045
[33] Elkins-Tanton, LT, Burgess, S, Yin, QZ. 2011. The lunar magma ocean: reconciling the solidification
process with lunar petrology and geochronology. Earth Planet. Sci. Lett. 304, 326-336.
[34] Piskorz, D, Stevenson, DJ. 2014. The formation of pure anorthosite on the Moon. Icarus 239, 238-243.
[35] Dygert, N, Lin, J-F, Marshall, EW, Kono, Y, Gardner, JE. 2017. A low viscosity lunar magma ocean forms a
stratified anorthitic flotation crust with mafic poor and rich units. Geophys. Res. Lett. 44, 11282-11291.
[36] Meyer J, Elkins-Tanton LT, Wisdom J. 2010. Coupled thermal-orbital evolution of the early Moon. Icarus
208, 1 -- 10
[37] Chen, EMA, Nimmo, F. 2013. Tidal dissipation in the lunar magma ocean and its effect on the early
evolution of the Earth-Moon system. Icarus 275, 132-142.
[38] Lin, Y, Tronche, EJ, Steenstra, ES, van Westrenen, W. 2016. Evidence for an early wet Moon from
experimental crystallization of the lunar magma ocean. Nat. Geosci. 10, 14-18.
[39] Lin, Y, Tronche, EJ, Steenstra, ES, van Westrenen, W. 2017. Experimental constraints on the solidification
of a nominally dry lunar magma ocean. Earth Planet. Sci. Lett. 471, 104-116.
[40] Hui, H, Peslier, AH, Zhang, Y, Neal, CR. 2013. Water in lunar anorthosites and evidence for a wet early
Moon. Nat. Geosci. 6, 177-180. (DOI: 10.1038/NGEO1735)
[41] Borg, LE, Norman, M, Nyquist, LE, Bogard, D, Snyder, GA, Taylor, L, Lindstrom, M. 1999. Isotopic
studies of ferroan anorthosite 62236: a young lunar crustal rock from a light rare-earth-element-depleted
source. Geochem. Cosmochim. Acta 63, 2679 -- 2691.
[42] Borg, LE, Shearer, CK, Nyquist, LE, Norman, M, 2002. Isotopic constraints on the origin of lunar ferroan
anorthosites. Proc. Lunar Planet. Sci. Conf. 33, 1396.
[43] Boyet, M., Carlson, RW. 2005. 142Nd evidence for early (>4.53 Ga) global differentiation of the silicate
Earth. Science 309, 576-581. (DOI: 10.1126/science.1113634)
Phil. Trans. R. Soc. A.
18
[44] Carlson RW, Lugmair GW 1988. The age of ferroan anorthosite 60025: oldest crust on a young Moon?
Earth Planet. Sci. Lett. 90, 119 -- 130.
[45] Nyquist LE, Wiesmann H, Bansal B, Shih C Y, Keith JE, Harper CL. 1995. 146Sm -- 142Nd formation interval
for the lunar mantle. Geochim. Cosmochim. Acta 59, 2817 -- 2837.
[46] Nyquist, LE, Shih, C-Y, Reese, YD, Park, J, Bogard, D, Garrison, D, Yamaguchi, A. 2010. Lunar crustal
history recorded in lunar anorthosites. Lunar Planet. Sci. Conf. 41, 1383.
[47] Kleine, T, Touboul, M, Bourdon, B, Nimmo, F, Mezger, K, Palme, H, Jacobsen, SB, Yin, Q-Z, Halliday, A.
2009. Hf-W chronology of the accretion and early evolution of asteroids and terrestrial planets. Geochim.
Cosmochim. Acta 73, 5150-5188.
[48] Borg, LE, Connelly, J, Boyet, M, Carlson, R, 2011. The age of lunar ferroan anorthosite 60025 with
implications for the interpretation of lunar chronology and the magma ocean model. Proc. Lunar and Planet.
Sci. Conf., 42, 1608.
[49] Boyet, M, Carlson, RW, Borg, LE, Horan, M. 2015. Sm-Nd systematics of lunar ferroan anorthositic suite
rocks: Constraints on lunar crust formation. Geochim. Cosmochim. Acta, 148, 203-218.
[50] Snape, JF, Nemchin, AA, Bellucci, JJ, Whitehouse, MJ, Tartèse, R, Barnes, JJ, Anand, M, Crawford, IA, Joy,
KH. 2016. Lunar basalt chronology, mantle differentiation and implications for determining the age of the
Moon. Earth Planet. Sci. Lett. 451, 149-158.
[51] Barboni, M, Boehnke, P, Keller, B, Kohl, IE, Schoene, B, Young, ED, McKeegan, KD. 2017. Early formation
of the Moon 4.51 billion years ago. Sci. Adv. 3, e1602365.
[52] Morbidelli, A, Nesvorny, D, Laurenz, V, Marchi, S, Rubie, DC, Elkins-Tanton, L, Wieczorek, M, Jacobson,
S. 2018. The timeline of the lunar bombardment: revisited. Icarus 305, 262-276.
[53] Hartung, JB. 1979. On the Global Asymmetry of the Lunar Crust Thickness. LPI Contribution 394, 45.
[54] Pernet-Fisher, J, Joy, K. 2016. The lunar highlands: old crust, new ideas. Astron. Geo. 57, 1.26-1.30.
[55] Blichert-Toft, J, Gleason, JD, Télouk, P, Albarède, F. 1999. The Lu-Hf isotope geochemistry of shergottites
and the evolution of the Martian mantle-crust system. Earth Planet. Sci. Lett. 173, 25-39.
[56] Borg, LE, Connelly, JN, Nyquist, LE, Shih, C-Y, Wiesmann, H, Reese, Y. 1999. The age of the carbonates in
martian meteorite ALH84001. Science 286, 90-94.
[57] Bouvier, LC, and 14 others. 2018. Evidence for extremely rapid magma ocean crystallization and crust
formation on Mars. Nature, in press.
[58] Dauphas, N. & Pourmand, A. 2011. Hf -- W -- Th evidence for rapid growth of Mars and its status as a
planetary embryo. Nature 473, 489 -- 492.
Phil. Trans. R. Soc. A.
19
[59] Senshu, H, Kuramoto, K, Matsui, T. 2002. Thermal evolution of a growing Mars. J. Geophys. Res. 107, 5118.
Doi:10.1029/2001JE001819.
[60] Ricard, Y, Sramek, O, Dubuffet, F. 2009. A multi-phase model of runaway core-mantle segregation in
planetary embryos. Earth Planet. Sci. Lett. 284, 144-150.
[61] Saito, H, Kuramoto, K. 2017. Formation of a hybrid-type proto-atmosphere on Mars accreting in the solar
nebula. Month. Not. Roy. Astron. Soc. 475, 1247-1287.
[62] Debaille, V, Brandon, AD, Yin, QZ, Jacobsen, B. 2007. Coupled 142Nd/143Nd evidence for a protracted
magma ocean in Mars. Nature 450, 525-528. Doi:10.1038/nature06317.
[63] Humayun, M, et al. 2013. Origin and age of the earliest Martian crust from meteorite NWA 7533. Nature
503, 513 -- 516. doi:10.1038/nature12764
[64] Nyquist, LE, Shih, CY, McCubbin, FM, Santos, AR, Shearer, CK, Peng, ZX, Burger, PV, Agee, CB. 2016.
Rb-Sr and Sm-Nd isotopic and REE studies of igneous components in the bulk matrix domain of Martian
breccia Northwest Africa 7034. Met. Planet. Sci. 51, 483-498.
[65] Elkins-Tanton, LT, Hess, PC, Parmentier, EM. 2005. Possible formation of ancient crust on Mars through
magma ocean processes. J. Geophys. Res. 110, E12S01. (doi:10.1029/2005JE002480)
[66] Elkins-Tanton, LT. 2008. Linked magma ocean solidification and atmospheric growth for Earth and Mars.
Earth Planet. Sci. Lett. 271, 181-191.
[67] Lebrun, T., Massol, H, Chassefiere, E, Davaille, A, Marcq, E, Sarda, P, Leblanc, F, Brandeis, G. 2013.
Thermal evolution of an early magma ocean in interaction with the atmosphere. J. Geophys. Res.: Planets 118,
1155-1176.
[68] Borg, LE, Brennecka, GA, Symes, SJK. 2016. Accretion timescale and impact history of Mars deduced from
the isotopic systematics of martian meteorites. Geochim. Cosmochim. Acta 175, 150-167.
[69] Bouvier, LC, et al. 2018. Evidence for extremely rapid magma ocean crystallization and crust formation on
Mars. Nature, 558, 586-589.
[70] Skok, JR, Mustard, JF, Tornabene, LL, Pan, C, Rogers, D, Murchie, SL. 2012. A spectroscopic analysis of
Martian crater central peaks: formation of the ancient crust. J. Geophys. Res. 117, E00J18. Doi:
10.1029/2012JE004148
[71] Plesa, AC, Tosi, N, Breuer, D 2014. Can a fractionally crystallized magma ocean explain the thermo-
chemical evolution of Mars? Earth Planet. Sci. Lett. 403, 225-235.
Phil. Trans. R. Soc. A.
20
[72] Ehlmann, BL, Edwards, CS. 2014. Mineralogy of the martian surface. Annu. Rev. Earth Planet. Sci. 42, 291-
315.
[73] Ramirez, R. 2017. A warmer and wetter solution for early Mars and the challenges with transient
warming. Icarus 297, 71-82.
[74] Wordsworth, R, Kalugina, Y, Lokshtanov, S, Vigasin, A, Ehlmann, B, Head, J, Sanders, C, Wang, H. 2013.
Transient reducing greenhouse warming on early Mars. Geophys. Res. Lett. 44, 665-671.
[75] Cannon, KM, Parman, SW, Mustard, JF. 2017. Primordial clays on Mars formed beneath a steam
supercritical atmosphere. Nature 552, 88-91. (doi:10.1038/nature24657)
[76] Sun, VZ, Milliken, RE. 2015. Ancient and recent clay formation on Mars as revealed from a global survey
of hydrous minerals in crater central peaks. J. Geophys. Res. Planets 120, 2293-2332.
[77] Dziewonski, AM, Hager, BH, O'Connell, RJ. 1977. Large-scale heterogeneities in the lower mantle. J.
Geophys. Res. 82, 239 -- 55
[78] Su, W-J, Woodward, RL, Dziewonski, AM, 1994.Degree 12 model of shear velocity heterogeneity in the
mantle. J. Geophys. Res. 99, 6945 -- 6980.
[79] Garnero EJ, McNamara AK. 2008. Structure and dynamics of Earth's lower mantle. Science 320, 626 -- 28
[80] Thorne MS, Garnero EJ. 2004. Inferences on ultralow-velocity zone structure from a global analysis of
SPdKS waves. J. Geophys. Res. 109, B08301
[81] Ishii, M, Tromp, J. 1999. Constraining large-scale mantle heterogeneity using mantle and inner-core
sensitive normal modes. Phys. Earth Planet. Int. 146, 113-124.
[82] Trampert, J, Deschamps, F, Resovsky, J, Yuen, D. 2004. Probabilistic tomography maps chemical
heterogeneities throughout the lower mantle. Science 306, 853-856
[83] Williams Q, Garnero EJ. 1996. Seismic evidence for partial melt at the base of Earth's mantle. Science 273,
1528 -- 30
[84] Lay, T, Garnero, EJ, Williams, Q. 2004. Partial melting in a thermo-chemical boundary layer at the base of
the mantle. Phys. Earth Planet. Inter. 146, 441 -- 67
[85] Rost, S, Garnero, EJ, Williams, Q, Manga, M. 2005. Seismological constraints on a possible plume root at
the core-mantle boundary. Nature 435, 666 -- 69
[86] McNamara AK, Garnero EJ, Rost S. 2010. Tracking deep mantle reservoirs with ultra-low velocity zones.
Earth Planet. Sci. Lett. 299, 1 -- 9
[87] Jellinek, AM, Manga, M 2002. The influence of a chemical boundary layer on the fixity, spacing and
lifetime of mantle plumes. Nature 418, 760-763.
Phil. Trans. R. Soc. A.
21
[88] McNamara, AK, Zhong, S. 2005. Thermochemical structures beneath Africa and the Pacific ocean. Nature
437, 1136-1139.
[89] Becker, TW, Kellogg, JB, O'Connell, R. 1999. Thermal constraints on the survival of primitive blobs in the
lower mantle. Earth Planet. Sci. Lett. 171, 351-365.
[90] Rizo, H, Walker, RJ, Carlson, RW, Touboul, M, Horan, MF, Puchtel, IS, Boyet, M, Rosing, MT. 2016. Early
Earth differentiation investigated through 142Nd, 182W, and highly siderophile element abundances in samples
from Isua, Greenland. Geochim. Cosmochim. Acta, 175, 319-336.
[91] Morino, P, Caro, G, Reisberg, L, Schumacher, A. 2017. Chemical stratification in the post-magma ocean
Earth inferred from coupled 146,147Sm-142,143Nd systematics in ultramafic rocks of the Saglek block (3.25-3.9 Ga;
northern Labrador, Canada). Earth Planet. Sci. Lett. 463, 136-150.
[92] Bouvier, A, Boyet, M 2016. Primitive Solar System materials and Earth share a common initial 142Nd
abundance. Nature 537, 399-402. (doi:10.1038/nature19351)
[93] Kurz, MD, Jenkins, WJ, Hart, SR. 1982. Helium isotopic systematics of oceanic islands and mantle
heterogeneity. Nature 297, 43-47.
[94] Kellogg, LH, Wasserburg, GJ 1990. The role of plumes in mantle helium fluxes. Earth Planet. Sci. Lett. 99,
276-289.
[95] Honda, R, Mizutani, H, Yamamoto, T. 1993. Numerical simulation of Earth's core formation. J. Geophys.
Res. 98, 2075-2089.
[96] Graham, DW 2002. Noble gas isotope geochemistry of midocean ridge and ocean island basalts:
characterization of mantle source reservoirs. Rev. Mineral. Geochem. 47, 247 -- 317.
[97] Tucker, JM, Mukhopadhyay, S. 2014. Evidence for multiple magma ocean outgassing and atmospheric
loss episodes from mantle noble gases. Earth Planet Sci. Lett. 393, 254-265.
[98] Hess, P, Parmentier, E, 1995. A model for the thermal and chemical evolution of the Moon's interior:
implications for the onset of mare volcanism. Earth Planet. Sci. Lett. 134, 501 -- 514.
[99] Ringwood, AE, Kesson, SE, 1976. A dynamic model for mare basalt petrogenesis. Proc. Lunar Sci. Conf. 7,
1697 -- 1722.
[100] Spera, FJ, 1992. Lunar magma transport phenomena. Geochem. Cosmochim. Acta 56,2253 -- 2265.
[101] Debaille, V, Brandon, AD, O'Neill, C, Yin, QZ, Jacobsen, B. 2009. Early martian mantle overturn inferred
from isotopic composition of nakhlite meteorites. Nat. Geosci. 2, 548-552. Doi:10.1038/ngeo579
Phil. Trans. R. Soc. A.
22
[102] Brown, SM, Elkins-Tanton, ET, Walker, RJ. 2014. Effects of magma ocean crystallization and overturn on
the development of 142Nd and 182W isotopic heterogeneities in the primordial mantle. Earth Planet. Sci. Lett.
408, 319-330. (https://doi.org/10.1016/j.epsl.2014.10.025)
[103] Labrosse, S, Hernlund, JW, Coltice, N. 2008. A crystallizing dense magma ocean at the base of the Earth's
mantle. Nature 450, 866-869. (doi:10.1038/nature06355)
[104] Lee, CA, Luffi, P, Höink, T, Li, J, Dasgupta, R, Hernlund, J. 2010 Upside-down differentiation and
generation of a 'primordial' lower mantle. Nature 463, 930-933. (doi:10.1038/nature08824)
[105] Tolstikhin, I, Hofmann, AW 2005. Early crust on top of the Earth's core. Phys. Earth Planet. Int. 148, 109-
130.
[106] Zhang, Z, Dorfman, SM, Labidi, J, Zhang, S, Li, M, Manga, M, Stixrude, L, McDonough, WF, Williams,
Q. 2016. Primordial metallic melt in the deep mantle. Geophys. Res. Lett. 43, 3693-3699.
(doi:10.1002/2016GL068560)
[107] Andrault, D, Bolfan-Casanova, N, Bouhifd, MA, Boujibar, A, Garbarino, G, Manthilake, G, Mezouar, M,
Monteux, J, Parisiades, P, Pesce, G. 2017. Toward a coherent model for the melting behaviour of the deep
Earth's mantle. Phys. Earth Planet. Int. 265, 67-81. (doi:10.1016/j.pepi.2017.02.009)
[108] Frost, DJ & McCammon, CA 2008. The redox state of Earth's mantle. Ann. Rev. Earth Planet. Sci. 36, 389-
420. Doi: 10.1146/annurev.earth.36.031207.124322
[109] Delano, JW 2001. Redox history of the Earth's interior since ~3900 Ma: implications for prebiotic
molecules. Origins Life Evol. Biosphere 31, 311-341.
[110] Canil, D 1997. Vanadium partitioning and the oxidation state of Archaean komatiite magmas. Nature,
389, 842-845.
[111] Li, ZXA & Lee, CTA 2004. The constancy of upper mantle fO2 through time inferred from V/Sc ratios in
basalts. Earth Planet. Sci. Lett. 228, 483-493.
[112] Nicklas, RW, Puchtel, IS, Ash, RD 2018. Redox state of the Archean mantle: Evidence from V partitioning
in 3.5-2.4 Ga komatiites. Geochim. Cosmochim. Acta 222, 447-466.
[113] Aulbach, S & Stagno, V 2016. Evidence for a reducing Archean ambient mantle and its effects on the
carbon cycle. Geology 44, 751-754.
[114] Trail, D, Watson, EB, Tailby, ND 2011.The oxidation state of Hadean magmas and implications for early
Earth's atmosphere. Nature 480, 79-82.
Phil. Trans. R. Soc. A.
23
[115] Badro, J., Brodholt, JP, Piet, H, Siebert, J, Ryerson, FJ. 2015. Core formation and core composition from
coupled geochemical and geophysical constraints. Proc. Nat. Acad. Sci. 112, 12310-12314.
Doi:10.1073/pnas.1505672112.
[116] Rubie, DC, Frost, DJ, Mann, U, Asahara, Y, Nimmo, F, Tsuno, K, Kegler, P, Holzheid, A, Palme, H 2011.
Heterogeneous accretion, composition and core-mantle differentiation of the Earth. Earth Planet. Sci. Lett. 301,
31-42.
[117] Rubie, DC, Jacobson, SA, Morbidelli, A, O'Brien, DP, Young, ED, de Vries, J, Nimmo, F, Palme, H, Frost,
DJ 2015 Accretion and differentiation of the terrestrial planets with implications for the compositions of early-
formed Solar System bodies and accretion of water. Icarus 248, 89-108.
[118] Sharp, ZD, McCubbin, FM, Shearer, CK 2013. A hydrogen-based oxidation mechanism relevant to
planetary formation. Earth Planet. Sci. Lett. 380, 88-97.
[119] Frost, DJ, Liebske, C, Langenhorst, F, McCammon, CA, Tronnes, RG, Rubie, DC 2004. Experimental
evidence for the existence of iron-rich metal in the Earth's lower mantle. Nature 428, 409-412.
[120] Wade, J, Wood, BJ 2005. Core formation and the oxidation state of the Earth. Earth Planet. Sci. Lett. 236,
78-95.
[121] Fischer, RA, Nakajima, Y, Campbell, AJ, Frost, DJ, Harries, D, Langenhorst, F, Miyajima, N, Pollock, K,
Rubie, DC 2015. High pressure metal-silicate partitioning of Ni, Co, V, Cr, Si, and O. Geochim. Cosmochim. Acta,
167, 177-194.
[122] Hirschmann, M 2012. Magma ocean influence on early atmosphere mass and composition. Earth Planet
Sci. Lett. 341-344, 48-57.
[123] Andrault, D, Bolfan-Casanova, N, Nigro, GL, Bouhifd, MA, Garbarino, G, Mezouar, M. 2011. Solidus and
liquidus profiles of chondritic mantle: implication for melting of the Earth across its history. Earth Planet. Sci.
Lett. 304, 251-259.
[124] Lange, RA, Carmichael, ISE 1989. Ferric-ferrous equilibria in Na2O-FeO-Fe2O3-SiO2 melts: effects of
analytical techniques on derived partial molar volumes. Geochim. Cosmochim. Acta 53, 2195-2204.
[125] Zhang, HL, Hirschmann, MM, Cottrell, E, Withers, AC 2017. Effect of pressure on Fe3+/ΣFe ratio in a
mafic magma and consequences for magma ocean redox gradients. Geochim. Cosmochim. Acta, 204, 83-103.
[126] Kress, VC, Carmicheal, ISE. 1991. The compressibility of silicate liquids containing Fe2O3 and the effect of
composition, temperature, oxygen fugacity and pressure on their redox states. Contrib. Mineral. Petrol. 108, 82-
92.
Phil. Trans. R. Soc. A.
24
[127] Komabayashi, T 2015. Thermodynamics of melting relations in the system Fe-FeO at high pressure:
implications for oxygen in the Earth's core. J. Geophys. Res.: Solid Earth 119, 4164-4177.
[128] Ricolleau, A, Fei, Y, Corgne, A, Siebert, J, Badro, J. 2011. Oxygen and silicon contents of Earth's core from
high pressure metal-silicate partitioning experiments. Earth Planet Sci. Lett. 310, 409-421.
[129] Armstrong, K, Frost, DJ, McCammon, CA, Rubie, DC, Boffa Ballaran, T. 2017. The effect of pressure on
iron speciation in silciate melts at a fixed oxygen fugacity: the possibility of a redox profile through a
terrestrial magma ocean. Am. Geophys. Un. 2017, abstract MR54A-04.
[130] Wadhwa, M. 2008. Redox conditions on small bodies, the Moon and Mars. Rev. Mineral. Geochem. 68, 493-
510.
[131] Zaranek, SE, Parmentier, EM 2004. The onset of convection in fluids with strongly temperature-
dependent viscosity cooled from above with implications for planetary lithospheres. Earth Planet. Sci. Lett. 224,
371-386.
[132] Scheinberg, A, Elkins-Taton, LT, Zhong, SJ. 2014. Timescale and morphology of Martian mantle overturn
immediately following magma ocean solidification. J. Geophys. Res. Planets 119, 454-467.
Doi:10.1002/2013JE004496
[133] Ballmer, MD, Lourenco, DL, Hirose, K, Caracas, R, Nomura, R. 2017. Reconciling magma-ocean
crystallization models with the present-day structure of the Earth's mantle. Geochem. Geophys. Geosys. 18, 2785-
2806. (doi:10.1002/2017GC006917)
[134] Maurice, M, Tosi, N, Samuel, H, Plesa, AC, Hüttig, C, Breuer, D. 2017. Onset of solid-state mantle
convection and mixing during magma ocean solidification. J. Geophys. Res. 122, 577-598.
(doi:10.1002/2016JE005250.)
[135] Tikoo, SM, Elkins-Tanton, LT. 2016. The fate of water within Earth and super-Earths and implications for
plate tectonics. Phil. Trans. Royal Soc. A 375, 20150394. (http://dx.doi.org/10.1098/rsta.2015.0394)
[136] Lenardic, A, Crowley, JW. 2012. On the notion of well-defined tectonic regimes for terrestrial planets in
this solar system and others. Astrophys. J. 755, 132.
[137] Weller, MB, Lenardic, A, O'Neill, C. 2015. The effects of internal heating and large scale climate
variations on tectonic bi-stability in terrestrial planets. Earth Planet. Sci. Lett. 420, 85-94.
[138] O'Neill, C, Lenardic, A, Weller, M, Moresi, L, Quenette, S, Zhang, S. 2016. A window for plate tectonics
in terrestrial planet evolution? Phys. Earth Planet. Inter. 255, 80-92.
[139] Nimmo, F, Stevenson, DJ 2000. Influence of early plate tectonics on the thermal evolution and magnetic
field of Mars. J. Geophys. Res. 105, 11969-11979.
Phil. Trans. R. Soc. A.
25
[140] Debaille, V, O'Neill, C, Brandon, AD, Haenecour, P, Yin, QZ, Mattielli, N, Treiman, A. 2013. Stagnant-lid
tectonics in early Earth revealed by 142Nd variations in late Archaen rocks. Earth Planet Sci Lett. 373, 83-92.
Doi:10.1016/j.epsl.2013.04.016
[141] Lourenco, DL, Rozel, AB, Gerya, T, Tackley, PJ. 2018. Efficient cooling of rocky planet by intrusive
magmatism. Nat. Geosci. 11, 322-327.
[142] Moore, WB, Simon, JI, Webb, AAG. 2017. Heat-pipe planets. Earth Planet. Sci. Lett. 474, 13-19.
[143] Foley, BJ, Bercovici, D, Elkins-Tanton, LT. 2014. Initiation of plate tectonics from post-magma ocean
thermochemical convection. J. Geophys. Res. Solid Earth 119, 8538-8561. (doi:10.1002/2014JB011121)
[144] O'Neill, C, Marchi, S, Zhang,S, Bootke, W. 2017. Impact-driven subduction on the Hadean Earth. Nat.
Geosci. 10, 793-797.
[145] Gerya, TV, Stern, RJ, Baes, M, Sobolev, SV, Whattam, SA. 2015. Plate tectonics on the Earth triggered by
plume-induced subduction initiation. Nature 527, 221-225.
[146] King, C, Olson, P. 2011. Heat partitioning in metal-silicate plumes during Earth differentiation. Earth
Planet. Sci. Lett. 304, 577-586.
[147] Fleck, JR, Rains, CL, Weeraratne, DS, Nguyen, CT, Brand, DM, Klein, SM, McGehee, JM, Rincon, JM,
Martinez, C, Olson, PL. 2018. Iron diapirs entrain silicates to the core and initiate thermochemical plumes. Nat.
Comm. 9, 71.
[125] Kowalski, M, Spencer, PJ 1995. Thermodynamic reevaluation of the Cr-O, Fe-O, and Ni-O systems:
remodelling of the liquid, bcc, and fcc phases. Calphad 19, 229-243.
[126] Matousek, J 2003. Thermodynamics of iron oxidation in metallurgical slags. J. Min. Metal Mater. Soc..64,
1314-1320.
Phil. Trans. R. Soc. A.
|
1306.2654 | 1 | 1306 | 2013-06-11T20:41:03 | Spatially-Resolved Millimeter-Wavelength Maps of Neptune | [
"astro-ph.EP"
] | We present maps of Neptune in and near the CO (2-1) rotation line at 230.538 GHz. These data, taken with the Combined Array for Research in Millimeter-wave Astronomy (CARMA) represent the first published spatially-resolved maps in the millimeter. At large (~5 GHz) offsets from the CO line center, the majority of the emission originates from depths of 1.1-4.7 bar. We observe a latitudinal gradient in the brightness temperature at these frequencies, increasing by 2-3 K from 40 degrees N to the south pole. This corresponds to a decrease in the gas opacity of about 30% near the south pole at altitudes below 1 bar, or a decrease of order a factor of 50 in the gas opacity at pressures greater than 4 bar. We look at three potential causes of the observed gradient: variations in the tropospheric methane abundance, variations in the H2S abundance, and deviations from equilibrium in the ortho/para ratio of hydrogen. At smaller offsets (0-213 MHz) from the center of the CO line, lower atmospheric pressures are probed, with contributions from mbar levels down to several bars. We find evidence of latitudinal variations at the 2-3% level in the CO line, which are consistent with the variations in zonal-mean temperature near the tropopause found by Conrath et al. (1998) and Orton et al. (2007). | astro-ph.EP | astro-ph |
Accepted to Icarus: 17 May 2013
Preprint typeset using LATEX style emulateapj v. 5/2/11
SPATIALLY-RESOLVED MILLIMETER-WAVELENGTH MAPS OF NEPTUNE
S. H. Luszcz-Cook12, I. de Pater13, M. Wright 1
Accepted to Icarus: 17 May 2013
ABSTRACT
We present maps of Neptune in and near the CO (2-1) rotation line at 230.538 GHz. These data,
taken with the Combined Array for Research in Millimeter-wave Astronomy (CARMA) represent
the first published spatially-resolved maps in the millimeter. At large (∼ 5 GHz) offsets from the
CO line center, the majority of the emission originates from depths of 1.1 -- 4.7 bar. We observe a
latitudinal gradient in the brightness temperature at these frequencies, increasing by 2 -- 3 K from
40◦N to the south pole. This corresponds to a decrease in the gas opacity of about 30% near the
south pole at altitudes below 1 bar, or a decrease of order a factor of 50 in the gas opacity at pressures
greater than 4 bar. We look at three potential causes of the observed gradient: variations in the
tropospheric methane abundance, variations in the H2S abundance, and deviations from equilibrium
in the ortho/para ratio of hydrogen. At smaller offsets (0 -- 213 MHz) from the center of the CO line,
lower atmospheric pressures are probed, with contributions from mbar levels down to several bars.
We find evidence of latitudinal variations at the 2-3% level in the CO line, which are consistent with
the variations in zonal-mean temperature near the tropopause found by Conrath et al. (1998) and
Orton et al. (2007).
NOTICE: this is the authors' version of a work that was accepted for publication in Icarus. Changes
resulting from the publishing process may not be reflected in this document.
1. INTRODUCTION
Neptune's millimeter continuum originates in the tro-
posphere, from pressures of 1 -- 5 bar. While collision-
induced absorption of H2 with hydrogen, helium and
methane dominates the opacity at these wavelengths,
several trace species also contribute, particularly H2S,
PH3 and NH3. The abundances of these trace species
have yet to be uniquely determined, though good fits
to centimeter-wavelength disk-integrated spectra, which
probe depths of several bars down to 10's of bars, are
obtained using an H2S abundance that represents a 30 --
50 times enrichment above the protosolar S/H value (de
Pater et al. 1991; DeBoer and Steffes 1996) and a pro-
tosolar abundance or less of nitrogen in NH3 (Romani
et al. 1989; de Pater et al. 1991).
Another trace atmospheric species, carbon monoxide
(CO), produces strong rotational lines at (sub)millimeter
wavelengths. CO is present in Neptune's upper atmo-
sphere at a level several orders of magnitude greater than
expected under thermochemical equilibrium conditions
(Marten et al. 1991). Two major pathways have been
identified for enriching Neptune's atmosphere in CO, and
these pathways result in different vertical CO abundance
profiles. The first of these is upward mixing of CO from
warm, deep layers of the atmosphere where CO is ther-
mochemically stable (Lodders and Fegley 1994); CO sup-
plied in this fashion will be well-mixed throughout the
upper atmosphere and therefore will exhibit a uniform
1 Astronomy Department, University of California, Berkeley,
CA 94720, USA
2 Astrophysics Department, American Museum of Natural
History, Central Park West at 79th Street, New York, NY 10024,
USA; [email protected]
3 SRON Netherlands Institute for Space Research, 3584 CA
Utrecht, and Faculty of Aerospace Engineering, Delft University
of Technology, 2629 HS Delft, The Netherlands
vertical profile. Alternatively, CO may be produced in
the stratosphere of the planet as a result of the infall of
oxygen-bearing material; in this case, downward trans-
port will act as a sink and the CO abundance will fall
below observable levels in the troposphere, where the dif-
fusion rate increases dramatically (Moses 1992).
To differentiate between these scenarios, models of
Neptune's disk-integrated vertical CO profile have been
produced using observations of the CO (1-0) (Luszcz-
Cook and de Pater 2013), (2-1) (Lellouch et al. 2005;
Luszcz-Cook and de Pater 2013) and (3-2) (Hesman et al.
2007) rotation lines at high spectral resolution (1.25-4
MHz) over a wide (8-20 GHz) frequency range. As illus-
trated in Fig. 1 for the CO (2-1) line at 230.538 GHz,
such observations are necessary to characterize the full
vertical CO profile, detecting emission from pressures be-
low 0.1 mbar at line center, up to several bars in the
far wings. From their respective studies, Lellouch et al.
(2005) and Hesman et al. (2007) find substantial tropo-
spheric CO abundances of 0.5±0.1 and 0.6±0.4 parts per
million (ppm). The analysis of Luszcz-Cook and de Pater
(2013), which favors a warmer temperature profile than
Lellouch et al. (2005) and Hesman et al. (2007), produces
a lower best-fit tropospheric CO abundance of 0.1+0.2−0.1
ppm. As first described by Prinn and Barshay (1977),
the observed tropospheric CO mole fraction represents
the equilibrium abundance at the CO 'quench level',
which is defined as the depth at which the timescale
for vertical mixing is equal to the timescale for chem-
ical conversion of CO into CH4. The equilibrium CO
mole fraction is directly proportional to the equilibrium
abundance of H2O, and under the conditions of Nep-
tune's deep atmosphere, nearly all the gas phase oxygen
is contained in water. Therefore, the CO abundance of
tropospheric CO acts as a probe of Neptune's global oxy-
2
S. H. Luszcz-Cook, I. de Pater, M. Wright
Fig. 1. -- Contribution functions for the CO (2 -- 1) line, illustrating the altitudes contributing to the observed intensity for a selection
of offsets from 0 to 5 GHz from line center. Contribution functions have been produced from the radiative transfer model, for disk center
(selected to be where the viewing angle µ (the cosine of the emission angle) is greater than 0.9, shown left), and near the limb (µ < 0.45,
right). The model was produced assuming a CO profile with 1.1 ppm CO in the stratosphere (at pressure less than 0.16 bar) and no CO in
the troposphere. CO opacity is greatest near line center; and emission at this frequency comes from the highest altitudes in the atmosphere
(black line). The cutoff in the CO abundance is responsible for the sharp decrease in the contribution functions near the tropopause.
Neptune's thermal profile has been plotted for reference (grey, see Section 3 for more information).
gen abundance. Luszcz-Cook and de Pater (2013) find
that a tropospheric CO mole fraction of 0.1 ppm implies a
global oxygen enrichment of at least 400, and likely more
than 650 times the protosolar O/H value. Note, though,
that the Luszcz-Cook and de Pater (2013) data are also
consistent with 0.0 ppm of CO in the troposphere, in
which case they do not constrain the global oxygen abun-
dance.
In addition to the CO abundance measured in
the troposphere, Lellouch et al. (2005), Hesman et al.
(2007), and Luszcz-Cook and de Pater (2013) all find
that the CO line shape is best fit by a CO abundance
profile that increases in the stratosphere, which suggests
that infall must also contribute to Neptune's observed
CO abundance. Based on the atmospheric CO/H2O ra-
tio, Lellouch et al. (2005) proposed that a recent large
cometary impact could be responsible for Neptune's ob-
served stratospheric CO abundance; however, comets of
the necessary size are expected to be exceedingly uncom-
mon. Luszcz-Cook and de Pater (2013) find that a con-
stant influx of (sub)kilometer-sized comets could supply
the observed stratospheric abundance of CO.
Spatially resolved maps of Neptune at centimeter
wavelengths have been obtained by several authors (de
Pater et al. 1991; Martin et al. 2006, 2008; Hofstadter
et al. 2008). Martin et al. (2006, 2008) and Hofstadter
et al. (2008) find a substantial (tens of K) increase in
the 1.3-2 cm brightness temperature near the south pole.
Such a temperature enhancement would result if dry air
subsides at this location, which would cause a decrease in
the atmospheric gas opacity so that warmer, deeper lay-
ers of the planet are probed. This observation is therefore
consistent with a global circulation pattern in which air
rises at mid- southern and northern latitudes and sub-
sides near the equator and south pole. Recently, Butler
et al. (2012) presented maps at 1 cm obtained with the
upgraded VLA, with a resolution of better than 0.1":
they observe that Neptune's bright polar cap extends
from the pole to 70◦S. They also see evidence for equa-
torial brightening, which would be consistent with the
circulation pattern outlined above.
In this paper, we
present the first spatially-resolved measurements of Nep-
tune at millimeter wavelengths, originally reported by
Luszcz-Cook et al. (2010). Our data set spans a range
of frequencies, from the center of the CO (2-1) line at
230.538 GHz (1.3 mm) to a maximum offset of 6 GHz
from line center, where continuum emission is detected.
The motivation behind these observations was to look
for latitudinal variations in the CO abundance, which
would provide additional information about the pattern
of CO infall/production. This was observed in the case
of Jupiter after the impact of comet Shoemaker-Levy 9
(SL9); Moreno et al. (2003) estimated that for SL9, lati-
0.00.20.40.60.81.0Contribution function10210010−210−410−6Pressure (bar)disk center050100150200250300Temperature (K)center5 MHz10 MHz20 MHz70 MHz200 MHz1 GHz5 GHzcenter5 MHz10 MHz20 MHz70 MHz200 MHz1 GHz5 GHz0.00.20.40.60.81.0Contribution function10210010−210−410−6Pressure (bar)limb050100150200250300Temperature (K)3
Fig. 2. -- Spectral coverage of our observations; for reference the model disk-integrated spectrum is shown in black. The short baseline
D-array data windows are indicated in red. The long baseline B-array bands are shown in pink (prior to 2011) and cyan (2011). For each
correlator setup, the narrow band is shown as a thicker line than the wide bands. As described in Section 2, our analysis focuses on the
frequencies where we have the most data, which are those spanned by the B-array data prior to 2011 (pink). The plot on the right is
zoomed in on line center.
tudinal variations in the CO abundance would persist for
roughly a decade. Furthermore, the CO abundance at a
given latitude depends on the rate of vertical and merid-
ional mixing, and could therefore act as a tracer of the
large-scale circulation. The spectrum in the CO line is
also affected by the temperature profile; therefore maps
in the CO line will be affected by horizontal variations
in temperature. In the 1.3 mm continuum, the intensity
depends on the gas opacity at depths of 1.1-4.7 bar which
may vary with latitude due to the large-scale circulation
pattern.
2. OBSERVATIONS
We observed Neptune with the Combined Array for
Research in Millimeter-wave Astronomy (CARMA), lo-
cated in the Inyo Mountains of eastern California 4.
CARMA consists of eight 3.5-meter antennas, nine 6-
meter antennas and six 10-meter antennas; our obser-
vations were performed with the 6- and 10-meter an-
tennas only, for a total of 105 baselines and 725 m2 of
collecting area. CARMA can be configured in 5 stan-
dard patterns 'A -- E', with the 'A' configuration being
the most extended and therefore producing the highest
angular resolution; and the 'E' configuration being the
most compact (lowest resolution). In order to spatially
resolve Neptune's 2.2" disk yet still 'see' the object and
have enough signal per synthesized beam, we observed
in the 'B' configuration, which has baselines of 82-946
meters and a synthesized beam of approximately 0.35"
at 230 GHz. Projected antenna separations ranged from
30 to 670 kλ.
Observations were carried out over a total of 13 days:
one test observing block, or 'track', was taken in De-
cember 2008, followed by eight tracks in the winter of
2009-2010 and four in January 2011. Of these, two of
the datasets from 2011 were of poor quality and were not
included in the analysis. Each dataset consists of a 15-
minute observation of the bright quasar 3C454.3 for pass-
band calibration, followed by a series of observing cycles
of 8 minutes on Neptune and 2 minutes (prior to 2011)
or 3 minutes (in 2011) on a nearby quasar (2229-085 in
December 2008 and January 2011; 3C446 in winter 2009-
2010) to be used for calibration of the atmospheric and
instrumental gains 5. The weather conditions during the
observations were generally fair, with root mean square
(rms) path errors ranging from 100 to 325 µm on a 100-
m baseline, and an average zenith optical depth of 0.18.
The total time on source in the B-array configuration
was 28 hours. To provide information at shorter base-
lines, we combined our B-array data with 15 hours on
source in the more compact D-array configuration from
spring 2009; these data have projected antenna separa-
tions of 5 -- 84 kλ and are described in Luszcz-Cook and
de Pater (2013).
Prior to upgrades in 2010, the CARMA correlator of-
fered three dual bands, or spectral windows, with config-
urable width of 500, 62, 31, 8 or 2 MHz. Each band could
be placed independently anywhere within the 4 GHz IF
bandwidth, and appears symmetrically in the upper and
lower sidebands of the first local oscillator. For our B-
array observations, we configured one correlator band to
62-MHz bandwidth and centered on the CO (2 -- 1) line at
230.538 GHz; the channel spacing in this 'narrow' band
was 0.9766 MHz. The remaining two bands were con-
figured to maximum (500 MHz) bandwidth and placed
at offsets from line center; the channel spacing in each
of these 'wide' bands was 31.25 MHz. The configuration
of the correlator during our 2008-2010 B-array observa-
tions is described in Table 1. Due to upgrades to the
4 CARMA is located at Cedar Flat, CA, at elevation 2200 m;
latitude 37.3; longitude -118.1
5 phase calibrators were typically located at a separation of 10 --
15 deg from Neptune. For more information on these calibrators,
see http://carma.astro.umd.edu/cgi-bin/calfind.cgi
220225230235frequency (GHz)810121416Flux density (Jy)B array, new correlatorB array, old correlatorD array229.54230.19230.84frequency (GHz)810121416Flux density (Jy)4
S. H. Luszcz-Cook, I. de Pater, M. Wright
B-array correlator setup, prior to 2011. This setup corresponds to configuration 'd' as specified in Table 1 of Luszcz-Cook and de Pater (2013). In
this paper only data within the frequencies spanned by this correlator setup are considered.
TABLE 1
Band center frequency Nominal bandwidtha Final bandwidthb Start frequencyb,c End frequencyb,c Channel spacing
224.387 GHz
225.537 GHz
229.538 GHz
230.688 GHz
230.538 GHz
224.247 GHz
225.397 GHz
229.397 GHz
230.547 GHz
230.517 GHz
281 MHz
281 MHz
281 MHz
281 MHz
42.0 MHz
500 MHz
500 MHz
500 MHz
500 MHz
62 MHz
224.528 GHz
225.678 GHz
229.679 GHz
230.829 GHz
230.559 GHz
31.25 MHz
31.25 MHz
31.25 MHz
31.25 MHz
0.9766 MHz
aCARMA correlator bands labeled as having a nominal width of 500 MHz bands actually have a total width of 468.75 MHz
bafter flagging of edge channels
cfrequency of band edge, NOT frequency at the center of the edge channel
correlator in late 2010, the channel spacing decreased in
both the 62-MHz and 500-MHz bands, and five addi-
tional dual bands were available during our 2011 obser-
vations. These additional bands were all configured to
maximum bandwidth mode, allowing for more continu-
ous coverage of the wide CO line. Figure 2 illustrates the
location of the correlator bands for the B-array observa-
tions prior to and after the correlator upgrade, as well as
the frequency coverage of the D-array data. In this pa-
per, we restrict our analysis to the frequencies spanned
by the original correlator setup (Table 1), where we have
the most data and therefore the greatest sensitivity.
We perform editing and calibration on each raw visi-
bility dataset using the Multichannel Image Reconstruc-
tion, Image Analysis and Display (MIRIAD; Sault et al.
2011) software package. After flagging edge channels
and poor quality data, we perform passband calibra-
tion. Then we correct for atmospheric and instrumen-
tal effects on the observed visibilities. We determine the
time-dependent antenna-based gains by self-calibrating
the wide-band data and then applying the gain solutions
to the full dataset. The absolute flux scale of each dataset
is set by scaling the antenna gains for each day of B-array
data so that the amplitudes of the visibilities match the
amplitudes of the D-array data at overlapping (u,v) dis-
tances6. A complete description of the data reduction is
provided in Appendix A.
3. MODEL
A series of model image cubes were created using the
line-by-line radiative transfer code described in Luszcz-
Cook and de Pater (2013), and using Neptune's atmo-
spheric properties as described therein. We adopt the
thermal profile of Fletcher et al. (2010) in the upper at-
mosphere, and an H2S abundance corresponding to 50
times the protosolar S/H abundance. As our nominal
case, we model the CO (2-1) absorption assuming that
the CO abundance is vertically distributed according to
the best-fit solution of Luszcz-Cook and de Pater (2013)
for the above-mentioned atmosphere 7: 1.1 ppm of CO
at altitudes above 0.16 bar and 0.0 ppm of CO deeper
than this level. At each of 155,575 locations on the disk
(0.005" pixels), we integrate the equation of radiative
transfer for the appropriate viewing angle µ (the cosine
6 The u and v coordinates describe the East-West and North-
South components of the projected interferometer baseline. The
(u,v) distance is defined as the projected baseline length.
7 Several of the atmospheric models in Luszcz-Cook and de Pater
(2013) do not include any H2S; here, we adopt the best-fit solution
for the model which has a 50 times solar abundance of H2S.
of the emission angle), accounting for the Doppler shift
due to the planet's rotation (Moreno et al. 2001). These
high-resolution models can be converted from brightness
temperature units into Jy/pixel, to be used as starting
models for the deconvolution process (Section 4, Ap-
pendix B). They can also be rebinned to the coarser
map resolution and convolved with a Gaussian beam,
and directly compared with the data (Section 5). For
the latter purpose, alternative models are also produced
by assuming the same vertical CO structure (the same
transition pressure, and no tropospheric CO), but vary-
ing the stratospheric CO abundance; or alternatively by
maintaining the nominal CO structure and abundance
but varying the temperature or composition (other than
CO) with latitude.
4. IMAGING
After combining all tracks of B- and D-array visibility
data, we produce images of Neptune in four steps: first
we use our visibility data to make a set of maps of the
estimated sky intensity distribution. The FWHM (full
width at half maximum) of the image point spread func-
tion, or synthesized beam, is in the range of 0.33 -- 0.37" x
0.37 -- 0.39", depending on frequency; the pixel size used
in the maps is 0.09". Next, we estimate the noise in
our maps due to calibration errors and atmospheric fluc-
tuations. We find that the noise level is approximately
three times greater than one would expect from thermal
noise from the atmosphere and receiver system, likely
because of the unfavorable weather during observations.
We then perform deconvolution to remove the response
to source structure in the side lobes of the synthesized
TABLE 2
Characteristics of final maps. Noise in each final map is calculated as
the root mean square of the noise in the input maps, divided by the
square root of the number of input images. See Appendix B.2 for a
discussion of noise estimation in the original (imaging) maps.
Average
frequency
offset from line center
0.0 MHz
4.6 MHz
9.2 MHz
13.8 MHz
18.4 MHz
66 MHz
213 MHz
1.00 GHz
5.58 GHz
frequency
Effective
widtha (MHz)
4.6
9.2
9.2
9.2
9.2
125
125
250
500
Noise in averaged
map (mJy/beam)
28.3
19.9
20.2
19.7
19.3
8.13
7.62
6.01
3.71
aTotal frequency width averaged to create the map.
beam. Finally, we average deconvolved images that are
at similar frequency offsets from the CO (2-1) line center,
in order to increase signal-to-noise without compromis-
ing our sensitivity to the vertical structure in Neptune's
atmosphere, i.e., we combine channels which have sim-
ilar contribution functions (Fig. 1). We also bin the
data according to latitude to look for meaningful latitu-
dinal trends in Neptune's observed brightness. Table 2
lists the central frequency and total frequency width of
each final image, along with our estimate of the noise in
units of mJy/beam. The details of the imaging process
are presented in Appendix B; this description includes
a discussion of the final error estimates and the binning
technique. A comparison of deconvolution methods is
given in Appendix C.
The images and latitude-binned plots are presented in
Figs. 3 -- 5. The top panel within each column is the fi-
nal, averaged map at the indicated frequency offset (δν)
from line center. The FWHM of the center of the synthe-
sized beam, which denotes the spatial resolution of the
map, is indicated by the filled red ellipse in the lower left
corner. The arrow points in the direction of Neptune's
north pole. Below each map, we also present two 'resid-
ual maps', which illustrate the variations between the
data and a 'flat' (uniform brightness) disk; and from the
nominal, horizontally uniform model described in Section
3 (second and third from the top, respectively). The first
of these residual maps emphasizes the variations across
the disk due to changes in emission angle (limb bright-
ening/darkening); the latter highlights horizontal varia-
tions in the composition and/or thermal profile. Below
each set of images, we present the latitude-binned data,
as compared to the latitude-binned flat disk and model.
To look for spatial variations in the observed atmospheric
properties, we also show the residuals between the hor-
izontally uniform nominal model and the data. These
figures are discussed further in the following section.
5. RESULTS
5.1. Continuum variations
The image at an average offset of 5.6 GHz from line
center (Fig. 3) is representative of the millimeter contin-
uum. The most noticeable spatial variation we observe
in the data is significant limb-darkening, as is predicted
by the radiative transfer mode. We also observe arti-
facts induced by calibration errors and atmospheric fluc-
tuations, which create a noticeable 'ringing' pattern in
the images. Averaging the data with respect to latitude
mitigates this issue, but higher-quality observations are
necessary to produce images without these artifacts.
Additionally, the data in Fig. 3 show two trends in
latitude with respect to the nominal (horizontally con-
stant composition and temperature) model. The first is
an overall offset in the brightness between the contin-
uum data and nominal model. We compare the inte-
grated flux density of the data and model and find that
the difference is −0.41 ± 0.02 Jy, or −2.9 ± 0.1 K. This
3% continuum offset could be due to issues in calibration
as well as real deficiencies in our model of the contin-
uum; this is discussed more thoroughly in Luszcz-Cook
and de Pater (2013). Here we concentrate on the second
observed trend in the data: latitudinal variations in the
continuum brightness. We observe that Neptune's south
5
Fig. 3. -- Average of the four maps with frequency offsets from
line center of 5 -- 6 GHz. Emission at these frequencies is primarily
millimeter continuum. Top image is the final, deconvolved and av-
eraged map. The average beam is indicated by the red circle in the
bottom left corner of the image. Second image from the top shows
the same data with a (beam-convolved) uniform brightness disk
subtracted. The white circle indicates the location of the planet's
limb, and the red arrow indicates the direction of the rotation axis.
The third image from the top has the beam-convolved nominal
(horizontally uniform composition and temperature) model sub-
tracted. In the plot (bottom) we zonally average the image (red
points) and nominal model (black) as described in Section 3. For
reference, we also show the average of the beam-convolved uniform
brightness disk as a function of latitude (blue). The differences
between the nominal model and the data (and between the nomi-
nal model and a uniform brightness disk) are shown in the bottom
part of the plot.
pole appears to be brighter than mid- and northern lati-
tudes; this is highlighted in panel (a) of Fig. 6, which is
a scaled version of Fig. 3. To determine the significance
of this result, we calculate χ2:
M−1(cid:88)
χ2 =
δy2
m
σ2
meas,m
m=0
(1)
where M is the number of data points in the fit, the pa-
6
S. H. Luszcz-Cook, I. de Pater, M. Wright
Fig. 4. -- Same as Fig. 3, except for δν = 66, 213, and 1000 MHz.
rameter δym is the difference between the data and model
values for point m, and σmeas,m is the measurement er-
ror for point m. We use the latitude-binned data for this
calculation, so that δym is the difference between the av-
erage flux density of the data and the model in bin m (in
Jy/beam); and σmeas,m is 3.71 mJy/beam (the noise es-
timate in the map within a single beam), divided by the
square root of the number of beams in bin m, as shown
by the error bars in Fig. 6. Since points from neighbor-
ing latitude bins are correlated (the width of each bin is
roughly equal to one beam), we include only every other
point in the statistical fit, starting with the data point
nearest the south pole, such that M = 4. The model that
we use in calculating χ2 is the nominal model scaled by a
constant factor of 0.97 to best match the data. We then
compare the value of χ2 to the probability p of measur-
ing that value for a set of experiments with a degree of
freedom
DOF = M − N
(2)
where N is the number fit parameters (in this case,
N = 1). We find that for these data compared to the
scaled, horizontally uniform nominal model, the calcu-
lated χ2 = 8.8. For a fit of 3 degrees of freedom there is
a probability p = 0.03 of a χ2 value this high occurring
due to chance. We therefore conclude that real spatial
variations in Neptune's millimeter continuum are likely,
and we explore the mechanisms that could cause such
variations.
The contribution function at an offset of 5 GHz from
the CO line center peaks at a depth of 4 bar, with most
of the emission originating from depths of 1.1-4.7 bar.
We expect that variations at these frequencies are likely
due to opacity variations, as opposed to horizontal tem-
perature variations, as the temperature profile is likely
adiabatic in the troposphere. A decrease in gas opacity
near the south pole would mean that warmer, deeper lay-
ers are probed. To estimate the magnitude of change in
optical depth that is required produce the observed flux
densities, we first match the data by scaling the total
optical depth in the model at the relevant pressures. We
find that decrease of about 30% in the south at pressures
greater than 1 bar is sufficient to match the observed lati-
tudinal gradient. If we vary the opacity only at pressures
greater than 2 bar, we require the opacity to be lower by
a factor of 2 in the south (e.g. from the nominal opacity
at 40◦N to 0.5 times nominal near the south pole). If we
restrict the opacity variations to pressures greater than 3
7
Fig. 5. -- Same as Fig. 3, except for δν = 0 -- 18 MHz. These images average over the smallest frequency interval (4.6 MHz for δν = 0,
9.2 MHz for the others), and therefore have the highest level of noise, as indicated by the error bars in the plots.
bar, we require a factor of 5-10 decrease in the opacity in
the south and if variations are only at pressures greater
than 4 bar, the opacity must be of order 50 times lower
in the south than at 40◦N to match the observations.
Therefore, the total opacity change across the disk im-
plied by our data depends on the pressures at which the
opacity changes, which is determined by what opacity
source is responsible. We investigate how variations in
three atmospheric properties might cause the observed
trend: the H2S abundance, the CH4 abundance and the
H2 ortho/para ratio.
The dominant opacity sources in our model at pres-
sures of 1.1-4.7 bar are H2S opacity and H2 collision-
induced absorption (CIA). For an H2S abundance of 50
times solar, H2S opacity begins to dominate over CIA
at pressures greater than 3.3 bar. The precise value
and variation of the H2S abundance in Neptune's tro-
posphere is poorly constrained, though cm-wavelength
observations suggest it may vary with latitude accord-
ing to the meridional circulation pattern (Martin et al.
2008; Hofstadter et al. 2008; Butler et al. 2012). We find
that we can reproduce the latitudinal behavior observed
in our data by varying the H2S opacity by a factor of
∼30 from the south pole to 40◦N. Figure 6b shows the
expected residuals for a case where the H2S opacity has
the nominal value from 10◦N to 40◦N, 0.3 times the nom-
inal value from 40◦S to 10◦N, and 0.03 times the nominal
value between 90◦S and 40◦S. This model has also been
scaled by a factor of 0.95 to match the integrated flux of
the data.
A decrease or increase in CH4 can also affect the bright-
ness temperature, even though methane on its own does
not contribute directly to the opacity (de Pater and
Mitchell 1993). CH4 condensation can change the adi-
abatic profile; however, we adopt a dry adiabat in the
upper atmosphere (Luszcz-Cook and de Pater 2013), so
CH4 does not affect our models in this way. Although
the CH4 mole fraction in the troposphere is only 2.2%
(Baines et al. 1995), the collision-induced absorption co-
efficient for H2-CH4 pairs is roughly 20 times higher than
for H2-H2 pairs under the relevant conditions. As a re-
sult, H2-CH4 CIA accounts for as much as 35% of the
total optical depth in the 1-4 bar region. As for H2S,
we determine the magnitude of variations that are re-
quired to reproduce the observed intensity gradient: we
find that the south polar region must be depleted in tro-
8
S. H. Luszcz-Cook, I. de Pater, M. Wright
Fig. 6. -- Latitude variations in the continuum map. Panel (a) is adapted from Fig. 3, with a different scaling of the color bar and y plot
axis. The image is the δν = 5.6 GHz data with the beam-convolved, horizontally uniform composition and temperature model subtracted.
The plot shows the latitude-binned residuals of the data−model. The filled white circle in the bottom left corner represents the full width
at half maximum of the beam. The black circle indicates the location of Neptune's limb, and the black arrow indicates the direction of the
rotation axis. Panel (b) shows a (scaled) model in which the H2S opacity has the nominal value from 10◦N to 40◦N, 0.3 times the nominal
value from 40◦S to 10◦N, and 0.03 times the nominal value between 90◦S and 40◦S. The nominal model has been subtracted as in (a). The
plot beneath the image shows the same data, binned in latitude (black diamonds). The data points from (a) are in grey. Panels (c) and
(d) are the same as (b), with different models: (c) shows a model where the tropospheric CH4 mole fraction is 0.044 (twice the nominal
value of 0.022) from 10◦N to 40◦N, .022 from 40◦S to 10◦N, and 0.0055 (25% of nominal) between 90◦S and 40◦S, is shown; (d) shows a
(scaled) model where the equilibrium hydrogen fraction is 0.6 from 10◦N to 40◦N, 0.8 from 40◦S to 10◦N, and 1.0 between 90◦S and 40◦S.
pospheric CH4 by a factor of order 10 relative to 40◦N.
Figure 6c shows the expected residuals for a case where
the tropospheric CH4 mole fraction is 0.044 (twice the
nominal value of 0.022) from 10◦N to 40◦N, 0.022 from
40◦S to 10◦N, and 0.0055 (25% of nominal) between 90◦S
and 40◦S. The disk-averaged CH4 mole fraction for this
example is 0.021, which is close to the expected nominal
value. As in Fig. 6b, a scale factor (in this case, 0.97)
has been applied to the model to match the data.
assumes 'intermediate' hydrogen as the nominal case.
That is, the ortho and para states of hydrogen are in
equilibrium at the local temperature, but the specific
heat is near that of 'normal' hydrogen. This situation
is described in Massie and Hunten (1982). Fast verti-
cal mixing from the deep interior, however, could bring
the ortho/para ratio closer to the 3:1 ratio expected for
normal hydrogen. We investigate the effect of variations
in the fraction of para-H2 on the opacity, while assum-
ing the adiabatic profile for normal hydrogen everywhere.
We find that normal hydrogen has a higher opacity than
'equilibrium' hydrogen. The average intensity is 5-6 K
Finally, we consider variations in the H2 ortho/para ra-
tio as a source of continuum intensity variations. As in
Luszcz-Cook and de Pater (2013), our nominal model
lower when we assume normal hydrogen rather than equi-
librium hydrogen. This implies that variations in the or-
tho/para ratio are more than capable of causing intensity
variations of the magnitude observed in our data. We can
match the observed latitudinal variations by decreasing
the fraction of hydrogen in the equilibrium state from 1.0
at the south pole to 0.6 at 40◦N (Fig. 6d), and scaling
this model by a factor of 0.98.
5.2. Variations near the tropopause
The images at frequency offsets of 0 -- 1000 MHz sense
lower pressures than the continuum map, due to a con-
tribution from CO to the opacity. Therefore these maps
include information on altitudes near and above the
tropopause at ∼0.1 bar (Fig. 1). Since observations to-
wards the limb sense higher altitudes, we expect maps
further from line center to be limb-darkened, whereas at
line center, where the emission originates in the strato-
sphere, we expect limb-brightening. This is indicated in
Figs. 4 and 5, which show the data relative to the nomi-
nal radiative transfer model and to a uniform brightness
('flat') disk. We find that the limb-darkening can be ob-
served above the noise in maps at offsets of 14-1000 MHz;
the limb-brightening signal is marginally detected in the
middle map (flat disk subtracted) at line center (δν =
0). Latitudinal trends in the data are more difficult to
observe in these maps than in the continuum map, since
these data are averaged over smaller frequency windows
and therefore have lower signal-to-noise. This is partic-
ularly true for the images in Fig. 5 which are typically
averaged over only 9.2 MHz. We concentrate our anal-
ysis on the latitude-binned data at offsets of 0-213 MHz
from line center, as shown in Fig. 7. Interpretation of
the 1000-MHz data is complicated by substantial contri-
butions from continuum opacity sources; therefore these
data are excluded from further analysis.
We note that the plots in Fig. 7 show a consistent lat-
itudinal behavior: the south pole appears bright relative
to our nominal model, southern mid-latitudes are rela-
tively dark, and additional brightening occurs near the
equator before the intensity falls off again at northern
mid-latitudes. Variations at these frequencies could be
due to variations in the CO abundance, variations in the
temperature, or a combination of the two. We there-
fore compare these data to three sets of models with
the following properties: horizontally uniform CO and
temperature profiles; a meridionally varying CO profile
and horizontally uniform temperature profile; and a hor-
izontally uniform CO profile and meridionally varying
temperature profile. For each model of interest, we cal-
culate χ2 as in Section 5.1, using every other data point
to avoid correlations between data. We determine χ2 for
each frequency separately (DOF = 4), as well as for the
data from all seven mapping frequencies simultaneously
(DOF = 28). The results are presented in Table 3 and
discussed below. For a fit with 4 degrees of freedom, a
single measured value of χ2
4 has a probability p = 0.05 of
being greater than 9.5. For a fit with 28 degrees of free-
dom, χ2
28 > 41.3 with a probability p = 0.05. Therefore,
we reject fits with χ2
4 > 9.5 and χ2
28 > 41.3.
5.2.1. Horizontally uniform CO and temperature
We find that for the maps 0-18 MHz from line center,
the nominal model, which maintains a uniform abun-
9
dance of 1.1 ppm of CO in the stratosphere, is consis-
tent with the data. At larger offsets from line center (66
and 213 MHz) the nominal model is rejected; at these
wavelengths the total frequency width averaged to pro-
duce the images is greater, resulting in smaller error bars;
therefore the deviations are more statistically significant.
When we consider the full dataset, we find that the nom-
inal model has an acceptable value of χ2 of 32.6.
In addition to our nominal model, we produce mod-
els with 0.5, 0.8, 1.2 and 1.5 times the nominal CO
abundance, with no latitudinal variation. Again we find
that at large offsets from line center these models are
inconsistent with the data. Figure 7 shows that, for
these frequencies, and increase or decrease of more than
20% in the CO abundance produces models that are too
high/low to match the data at δν = 66 and 213 MHz,
for the given spatial resolution. At smaller offsets, the
effect of changing the CO abundance becomes more com-
plex. At an offset of 5 MHz, for example, an increase
or decrease in the CO abundance will lead to an over-
all brighter model: increasing the CO abundance would
increase the limb brightening, while decreasing the CO
abundance would increase the intensity at disk center.
We conclude that models of this form (with no latitu-
dinal variations in the CO abundance or temperature)
are not capable of reproducing the observed trend in the
CO abundance with latitude, and models with higher and
lower CO abundances significantly decrease the goodness
of fit of the full dataset.
5.2.2. Meridionally varying CO, horizontally uniform
temperature
An interesting possibility is that the observed varia-
tions in Neptune's brightness could be due to latitudinal
variations in the CO abundance, related to localized in-
fall or circulation effects. We compare the data to several
simple models. First, we increase the CO abundance by
20% over the nominal value in the northern hemisphere
and decrease it by 20% in the south, and vice versa. A
southern hemisphere increase in the CO abundance is
the better match of the two, but does not reproduce the
brightening we observe near the south pole and causes
an increase in χ2 for the 66 MHz, 213 MHz, and com-
bined datasets. We also try three slightly more complex
models. In the first case, we increase the CO abundance
by 20% at high latitudes (above 60◦), and decrease it
by 20% elsewhere. We repeat this for an increase at
mid-latitudes (between 30 and 60◦N and S) and near the
equator (30◦S to 30◦N). We find that an increase in the
CO at mid-latitudes provides a qualitatively good match
to the data at 66 and 213 MHz and decreases the value of
χ2 for the full dataset. However, for frequency offsets less
than 66 MHz, the quality of the fit generally decreases.
Despite this, we do not rule out a model in which the
CO abundance is higher at mid-latitudes.
5.2.3. Horizontally uniform CO, meridionally varying
temperature
Variations in the zonal mean temperature have been
observed near Neptune's tropopause in infrared images
from the Voyager 2 spacecraft (Conrath et al. 1998) and
the Very Large Telescope (VLT) (Orton et al. 2007). The
latitude coverage of the Voyager IRIS data extend from
80◦S to 30◦N: Conrath et al. (1998) find a 100-mbar
10
S. H. Luszcz-Cook, I. de Pater, M. Wright
Fig. 7. -- Comparison of the zonally-averaged data and models, for frequency offsets of 0 -- 213 MHz. First two columns are the CLEAN maps
and nominal-model subtracted maps. The full width at half maximum of the beam is shown by a filled red circle. In the difference maps, the
limb is indicated by a white circle and the direction of the rotation axis is shown by a red arrow. For each set of data−model comparisons,
the top portion of each plot shows the binned data (red squares) with errors and several comparison models (identified at the top of each
column). Below each plot is a difference plot, showing the same data and models with the nominal model subtracted, highlighting spatial
variations in brightness due to variations in opacity and/or temperature. The first column of plots shows a selection of horizontally uniform
CO, uniform temperature models (Section 5.2.1). The second column of plots presents horizontally uniform temperature, meridionally
varying CO models (Section 5.2.2). The final column presents three horizontally uniform CO, meridionally varying temperature models
(Section 5.2.3).
temperature minimum (4-5 K cooler than the equator)
near 45◦S, as well as a decrease in temperature at the
northernmost extent of their data. The measurements
of Orton et al. (2007) extend all the way to the south
pole. These authors find similar latitudinal behavior of
the temperature from their 17.6 (Orton model 'A') and
18.7 µm VLT (Orton model 'B') images. Near the south
pole, they observe 100-mbar temperatures that are 7-10
K higher than elsewhere on the planet.
To test the impact of these temperature variations on
our models, we adjust the Fletcher et al. (2010) tem-
perature profile with latitude to match the temperature
variations reported by Conrath et al. (1998) and Orton
et al. (2007) at 100 mbar. For simplicity, the tempera-
ture profile is modified by the same temperature offset
at all altitudes above 1 bar. We acknowledge that this
paradigm is most likely overly simplistic: the 2D tem-
perature cross sections of Conrath et al. (1998) show
that temperature variations appear to be located pri-
marily near the tropopause, with a gradual decrease in
the meridional variations with altitude to nearly uniform
within a few scale heights. However, Fig. 1 illustrates
that much of the emission at 5-200 MHz from line center
originates from near the tropopause, and from contin-
uum opacity sources well below the tropopause (where
the temperature should not vary and is not adjusted).
We expect that models of this form are adequate for a
first test of the effects of temperature variations on our
maps. We find that these varying temperature models
provide a good match to the data overall, and decrease
the value of χ2 at all frequencies (Fig. 7 and Table 3).
The Orton models of the zonal mean temperature pro-
file are in general a better match to the data than the
Conrath et al. (1998) solution; this is partially because
the Conrath et al. (1998) model does not extend to the
south pole.
The frequency at which the least improvement is ob-
served is 213 MHz from line center. Figure 7 shows that
the varying temperature models do qualitatively repro-
duce the latitudinal behavior seen at this frequency; how-
ever, there appears to be an overall offset between the
data and these models. Of the wavelengths considered,
the data at δν =213 MHz has the greatest contribution
from continuum emission (Fig. 1), and therefore we con-
sider the possibility that such an offset exists due to the
continuum issues discussed in Section 5.1. To test this,
we refit each of the models for δν = 213 MHz, allow-
ing for an overall scaling of the models to best match
the data. We find that the fit improves dramatically
for the varying temperature models when we allow for
such a scale factor. We note that the fit also improves
for several other models when we allow a scale factor in
the fitting process; however we caution that in several
instances the scale factors found in the fit have a value
greater than 3%, which seems unlikely given the results
of Section 5.1. These cases are indicated in Table 3.
To summarize, we find that the following models are
not rejected by our χ2 analysis: the nominal model, in
which the CO and thermal profiles are horizontally uni-
form; the model in which the stratospheric CO abun-
dance is increased at mid-latitudes and decreased else-
where; and the three models in which the temperature
profile varies with latitude to match the results of Or-
ton et al. (2007) and Conrath et al. (1998). Of these
models, the nominal model appears to be ruled out as a
fit to either the δν =66 MHz or the δν =213 MHz data
when considered alone; furthermore, the four indicated
test models all offer an improvement in the fit quality to
the full dataset over the nominal model, with the greatest
improvement for the varying temperature models. This
improvement is even more dramatic when we allow the
models at 213 MHz to have an applied scale factor to
account for continuum opacity issues.
6. SUMMARY AND CONCLUSIONS
Maps of Neptune are presented at frequency offsets of
δν = 0 MHz -- 5.6 GHz from the center of the CO (2-1)
line at 230.538 GHz. Far from the CO line, the observed
emission is primarily from the continuum, and the opac-
ity is dominated by H2S absorption and H2 CIA. Our
observations show that the region near the south pole
is brighter than other latitudes. Similarly, centimeter-
11
wavelength observations have shown that the south pole
(Martin et al. 2006, 2008; Hofstadter et al. 2008) and
equator (Butler et al. 2012) are bright compared to the
rest of the planet. This has been attributed to a global
circulation pattern in which air rises at mid- southern
and northern latitudes and subsides near the equator and
south pole; the subsidence of dry air would decrease the
gas opacity allowing thermal emission from deeper levels
to be observed. We model the potential effects of three
opacity sources on Neptune's millimeter continuum: the
H2S abundance, the tropospheric methane abundance,
and deviations from equilibrium in the ortho/para ratio
of hydrogen. We find that our observations can be re-
produced by latitudinal variations in the H2S opacity; a
good match to the data is found by a model in which the
H2S opacity has the nominal value from 10◦N to 40◦N,
0.3 times the nominal value from 40◦S to 10◦N, and 0.03
times the nominal value between 90◦S and 40◦S. Such a
decrease in the H2S opacity near the south pole could
cause an increase in the brightness temperature in this
location at longer wavelengths as well.
For opacity sources that dominate at higher altitudes
in the atmosphere, such as H2 CIA, a smaller change in
the opacity would be enough to produce the latitudinal
trend in brightness observed in our continuum data. We
can alternatively model the observed latitudinal varia-
tions by decreasing the CH4 mole fraction from 0.044 at
10◦N -- 40◦N to 0.022 at 40◦S -- 10◦N and 0.0055 be-
tween 90◦S and 40◦S. While most previous studies (e.g.
Roe et al. 2001; Gibbard et al. 2002; Irwin et al. 2011)
assume that Neptune's methane mole fraction is con-
stant in the troposphere with latitude, Karkoschka and
Tomasko (2011) find that CH4 is depressed between 1.2
and 3.3 bar at high southern latitudes, compared to its
abundance at low latitudes. While we estimate some-
what larger variations in the CH4 mole fraction (a to-
tal change of a factor of 8 rather than a factor of ∼3
found by Karkoschka and Tomasko (2011)), the similar-
ities between our estimated latitudinal CH4 profile and
that derived by Karkoschka and Tomasko (2011) are con-
spicuous. Variations in the ortho/para ratio of hydrogen
can also cause variations in the continuum intensity, and
we find that a decrease from an equilibrium hydrogen
fraction of 1.0 at the south pole to 0.6 at 40◦N produces
the observed latitudinal trend. Baines et al. (1995) deter-
mined that the hydrogen in Neptune's upper troposphere
is near equilibrium; however, Conrath et al. (1998) mea-
sured a decrease in the para hydrogen fraction from equi-
librium at latitudes of 0◦ -- 60◦S, and an increase in the
para fraction over the equilibrium value near the south
pole and in the northern hemisphere. This result qualita-
tively agrees with the pattern we observe near the south
pole and equator, although we do not see a rise in the
intensity in the northern hemisphere in our data, which
we would expect from an increase in the para fraction
at these latitudes. We note that we do not investigate
the effects of changing the adiabatic profile. This has
been studied previously: de Pater and Mitchell (1993)
show the effect of equilibrium, normal and intermediate
hydrogen on Neptune's microwave spectrum between 0.1
and 10 mm by changing both the adiabatic profile and
the CIA opacity. They show that the brightness tem-
perature is highest for intermediate H2, and lowest for
equilibrium H2.
12
S. H. Luszcz-Cook, I. de Pater, M. Wright
TABLE 3
Calculated χ2 values for the latitude-binned data and a set of test models. See Section 5.2 for a description of each model. Every other bin,
starting with the bin centered at 80◦S, is used in calculated χ2, since data from neighboring bins are correlated. For individual maps (unscaled
models), DOF = 4 and the probability p of obtaining χ2 > χ2
4 > 9.5. For the scaled models at δν = 213
4 due to chance is less than 0.05 for χ2
3 > 7.8. When considering all frequency offsets simultaneously, DOF = 28 and p ≤ 0.05 for χ2
28 > 41.3
MHz, DOF = 3 and p ≤ 0.05 for χ2
(DOF = 27 and p ≤ 0.05 for χ2
27 > 40.1 for all data when the δν = 213 MHz data are scaled.)
Frequency offset χ2
0.0 MHz
4.6 MHz
9.2 MHz
13.8 MHz
18.4 MHz
66 MHz
213 MHz
nominal
model
1.9
2.8
2.0
2.8
2.1
10.8
10.2
50% CO 80%
CO
0.9
3.9
3.9
7.6
7.3
33.0
15.3
2.9
10.4
15.8
30.0
33.5
206.5
145.1
213 MHz, scaled 6.3
all
all,
scaled
213 MHz
32.6
28.7
4.5a
444.2
303.6
5.7
71.9
62.3
120%
CO
5.9
2.8
1.5
1.4
1.1
29.6
49.7
6.7a
91.9
49.0
150%
CO
18.8
4.5
1.4
1.4
2.5
97.5
165.1
6.9a
291.3
133.1
+CO in
N
1.6
4.6
3.9
6.5
6.3
31.7
20.6
+CO in
S
4.3
2.2
1.3
1.7
0.9
14.6
25.3
+CO
high lat
2.0
3.7
3.9
7.6
7.0
34.0
19.4
19.4
75.2
74.0
5.3
50.3
30.3
14.2
77.5
72.2
+CO
mid-lat
2.9
3.1
2.1
4.2
2.7
3.0
1.5
1.2
19.5
19.2
+CO
equator
1.3
3.6
3.0
3.7
4.5
37.5
34.9
Orton
2007 'A'
1.2
1.2
0.9
1.9
0.9
4.3
8.9
Orton
2007 'B'
1.2
1.2
0.8
1.8
0.8
2.6
7.3
Conrath
1998
0.9
1.3
1.5
2.5
0.9
3.1
8.7
31.2
88.5
84.9
0.7
19.3
11.2
0.7
15.7
9.1
2.2
18.9
12.4
aBest-fit scale factor is greater than 3%, which is considered unlikely given the results for the continuum data.
which is located on the Chajnantor plateau at an altitude
of 5000m and will consist of 66 7- and 12-m antennas
when completed, will provide unprecedented sensitivity
and resolution for such observations.
ACKNOWLEDGEMENTS
The data presented in this work were obtained with
CARMA. Support for CARMA construction was derived
from the states of California, Illinois, and Maryland,
the James S. McDonnell Foundation, the Gordon and
Betty Moore Foundation, the Kenneth T. and Eileen L.
Norris Foundation, the University of Chicago, the As-
sociates of the California Institute of Technology, and
the National Science Foundation. Ongoing CARMA
development and operations are supported by the Na-
tional Science Foundation under a cooperative agree-
ment, and by the CARMA partner universities. This
work was supported by NASA Headquarters under the
NASA Earth and Space Science Fellowship program -
Grant NNX10AT17H; and by NSF Grant AST-0908575.
The authors would like to thank L. Fletcher for provid-
ing his temperature and CH4 profiles, and G. Orton for
providing his H2 CIA absorption coefficients. The au-
thors would also like to acknowledge R.L. Plambeck and
A. Bauermeister for many helpful discussions.
At frequency offsets δν =0 -- 213 MHz from line center,
where we are sensitive to the CO abundance and temper-
ature in the lower stratosphere, we find that the south
pole and equator are relatively bright compared to the
region near 45◦S and 40◦N. Maps at δν =66 MHz or the
δν =213 MHz are inconsistent with the nominal model,
which has a horizontally uniform CO and thermal pro-
file. Introducing the temperature variations derived by
Conrath et al. (1998) or Orton et al. (2007) into our mod-
els successfully reproduces the observed latitudinal vari-
ations. As discussed by Conrath et al. (1998) and Martin
et al. (2008), the low temperatures at mid-latitudes are
consistent with upward motion and adiabatic cooling, as
suggested by the centimeter and millimeter continuum
data. The increase in temperature near the south pole
is likely due to increased solar insolence in this region
(Hammel et al. 2007; Orton et al. 2007) and/or subsi-
dence and adiabatic heating (Martin et al. 2008). We also
compare the data to models in which the CO abundance
varies with latitude; we find that a model in which the
stratospheric CO abundance is increased at mid-latitudes
and decreased elsewhere is also consistent with the data.
We expect that any variations in the CO abundance must
be less than 20%, at the resolution of our observations.
These data suggest the utility of measurements in and
near CO rotation lines for constraining latitudinal vari-
ations in the thermal profile and opacity in Neptune's
lower stratosphere and upper troposphere. Future ob-
servations can improve on these results in several ways.
First of all, poor weather during our observations in-
creased the noise level and introduced ripples into the
final maps, limiting the power of the dataset for ruling
out alternative models. Secondly, in addition to being
interesting in its own right, a better characterization of
Neptune's continuum opacity is critical to better model-
ing data within the CO line. For example, our analysis
from Luszcz-Cook and de Pater (2013) indicates that H2S
opacity has a small but non-negligible effect on the re-
trieved vertical CO profile. Finally, in order to disentan-
gle the effect of temperature and CO variations, spatial
maps of several CO transitions are required. The At-
acama Large Millimeter/sub-millimeter Array (ALMA),
REFERENCES
13
Baines, K. H., Mickelson, M. E., Larson, L. E., and Ferguson,
D. W. (1995). The abundances of methane and ortho/para
hydrogen on Uranus and Neptune: Implications of New
Laboratory 4-0 H2 quadrupole line parameters. Icarus,
114:328 -- 340.
Briggs, D. S. (1995). High Fidelity Interferometric Imaging:
Robust Weighting and NNLS Deconvolution. In American
Astronomical Society Meeting Abstracts, volume 27 of Bulletin
of the American Astronomical Society, page 112.02.
Butler, B. J., Hofstadter, M., Gurwell, M., Orton, G., and
Norwood, J. (2012). The Deep Atmosphere of Neptune From
EVLA Observations. In AAS/Division for Planetary Sciences
Meeting Abstracts, volume 44 of AAS/Division for Planetary
Sciences Meeting Abstracts, page 504.06.
Clark, B. G. (1980). An efficient implementation of the algorithm
'CLEAN'. A&A, 89:377.
Conrath, B. J., Gierasch, P. J., and Ustinov, E. A. (1998).
Thermal Structure and Para Hydrogen Fraction on the Outer
Planets from Voyager IRIS Measurements. Icarus, 135:501 -- 517.
de Pater, I. and Mitchell, D. L. (1993). Radio observations of the
planets - The importance of laboratory measurements.
J. Geophys. Res., 98:5471 -- 5490.
de Pater, I., Romani, P. N., and Atreya, S. K. (1991). Possible
microwave absorption by H2S gas in Uranus' and Neptune's
atmospheres. Icarus, 91:220 -- 233.
DeBoer, D. R. and Steffes, P. G. (1996). Estimates of the
Tropospheric Vertical Structure of Neptune Based on
Microwave Radiative Transfer Studies. Icarus, 123:324 -- 335.
Fletcher, L. N., Drossart, P., Burgdorf, M., Orton, G. S., and
Encrenaz, T. (2010). Neptune's atmospheric composition from
AKARI infrared spectroscopy. A&A, 514:A17.
Gibbard, S. G., Roe, H., de Pater, I., Macintosh, B., Gavel, D.,
Max, C. E., Baines, K. H., and Ghez, A. (2002).
High-Resolution Infrared Imaging of Neptune from the Keck
Telescope. Icarus, 156:1 -- 15.
Hammel, H. B., Sitko, M. L., Lynch, D. K., Orton, G. S., Russell,
R. W., Geballe, T. R., and de Pater, I. (2007). Distribution of
Ethane and Methane Emission on Neptune. AJ, 134:637 -- 641.
Hesman, B. E., Davis, G. R., Matthews, H. E., and Orton, G. S.
(2007). The abundance profile of CO in Neptune's atmosphere.
Icarus, 186:342 -- 353.
Hofstadter, M. D., Butler, B. J., Gurwell, M. A., Hesman, B. E.,
and Devaraj, K. (2008). The Tropospheres of Uranus and
Neptune as seen at Microwave Wavelengths. In AAS/Division
for Planetary Sciences Meeting Abstracts #40, volume 40 of
Bulletin of the American Astronomical Society, page 488.
Irwin, P. G. J., Teanby, N. A., Davis, G. R., Fletcher, L. N.,
Orton, G. S., Tice, D., Hurley, J., and Calcutt, S. B. (2011).
Multispectral imaging observations of Neptune's cloud
structure with Gemini-North. Icarus, 216:141 -- 158.
Karkoschka, E. and Tomasko, M. G. (2011). The haze and
methane distributions on Neptune from HST-STIS
spectroscopy. Icarus, 211:780 -- 797.
Lellouch, E., Moreno, R., and Paubert, G. (2005). A dual origin
for Neptune's carbon monoxide? A&A, 430:L37 -- L40.
Lodders, K. and Fegley, Jr., B. (1994). The origin of carbon
monoxide in Neptune's atmosphere. Icarus, 112:368 -- 375.
Luszcz-Cook, S. H. and de Pater, I. (2013). Constraining the
origins of Neptune's carbon monoxide abundance with
CARMA millimeter-wave observations. Icarus, 222:379 -- 400.
Luszcz-Cook, S. H., de Pater, I., and Plambeck, R. L. (2010).
Spatially Resolved Observations of CO in the Atmosphere of
Neptune. In AAS/Division for Planetary Sciences Meeting
Abstracts #42, volume 42 of AAS/Division for Planetary
Sciences Meeting Abstracts, page 11.33.
Marten, A., Gautier, D., Owen, T., Sanders, D., Tilanus, R. T.,
Deane, J., and Matthews, H. (1991). First Detections of CO
and HCN in the Atmosphere of Neptune. In Bulletin of the
American Astronomical Society, volume 23 of Bulletin of the
American Astronomical Society, page 1164.
Martin, S., de Pater, I., and Kloosterman, J. andHammel, H. B.
(2008). Multi-wavelength observations of neptune's
atmosphere. In EPSC Abstracts, pages EPSC2008 -- A -- 00277.
Martin, S., de Pater, I., Kloosterman, J., Gibbard, S., and
Hammel, H. B. (2006). Multi Wavelength imaging of Neptune
at High Spatial Resolution. In AAS/Division for Planetary
Sciences Meeting Abstracts #38, volume 38 of Bulletin of the
American Astronomical Society, page 502.
Massie, S. T. and Hunten, D. M. (1982). Conversion of para and
ortho hydrogen in the Jovian planets. Icarus, 49:213 -- 226.
Moreno, R., Marten, A., Biraud, Y., B´ezard, B., Lellouch, E.,
Paubert, G., and Wild, W. (2001). Jovian stratospheric
temperature during the two months following the impacts of
comet Shoemaker-Levy 9. Planet. Space Sci., 49:473 -- 486.
Moreno, R., Marten, A., Matthews, H. E., and Biraud, Y. (2003).
Long-term evolution of CO, CS and HCN in Jupiter after the
impacts of comet Shoemaker-Levy 9. Planet. Space Sci.,
51:591 -- 611.
Moses, J. I. (1992). Meteoroid ablation in Neptune's atmosphere.
Icarus, 99:368 -- 383.
Orton, G. S., Encrenaz, T., Leyrat, C., Puetter, R., and Friedson,
A. J. (2007). Evidence for methane escape and strong seasonal
and dynamical perturbations of Neptune's atmospheric
temperatures. A&A, 473:L5 -- L8.
Prinn, R. G. and Barshay, S. S. (1977). Carbon monoxide on
Jupiter and implications for atmospheric convection. Science,
198:1031 -- 1034.
Rich, J. W., de Blok, W. J. G., Cornwell, T. J., Brinks, E.,
Walter, F., Bagetakos, I., and Kennicutt, Jr., R. C. (2008).
Multi-Scale CLEAN: A Comparison of its Performance Against
Classical CLEAN on Galaxies Using THINGS. AJ,
136:2897 -- 2920.
Roe, H. G., Gavel, D., Max, C., de Pater, I., Gibbard, S.,
Macintosh, B., and Baines, K. H. (2001). Near-Infrared
Observations of Neptune's Tropospheric Cloud Layer with the
Lick Observatory Adaptive Optics System. AJ, 122:1636 -- 1643.
Romani, P. N., de Pater, I., and Atreya, S. K. (1989). Neptune's
deep atmosphere revealed. Geophys. Res. Lett., 16:933 -- 936.
Sault, R. J., Teuben, P. J., and Wright, M. C. H. (2011).
MIRIAD: Multi-channel Image Reconstruction, Image
Analysis, and Display. In Astrophysics Source Code Library,
record ascl:1106.007, page 6007.
APPENDIX
DATA REDUCTION
We perform editing and calibration on the raw visibility data using the Multichannel Image Reconstruction, Image Analysis and Display
(MIRIAD; Sault et al. 2011) software package. We flag 10 edge channels in the narrow band centered on the CO line and three edge channels
in the wide bands; the narrow window in the opposite sideband from the CO line is flagged entirely. Poor quality data are also flagged.
After performing passband calibration, we correct for atmospheric and instrumental effects on the observed visibilities by determining
the time-dependent antenna-based gains using the wide-band data. An initial self calibration is performed in two steps using the phase
calibrator: a record-by-record phase solution is found for the phase calibrator to remove short-term variations. This is followed by a phase
and amplitude self calibration on the phase calibrator using a time interval corresponding to one observing cycle; these gain solutions are
then applied to the Neptune data to correct for slow variations in the antenna gains. Finally, a record-by-record phase-only self calibration
is performed on Neptune itself to remove short-term phase variations in the Neptune data. We find that since the rms path errors are a
large fraction of the wavelength of the observations, the phase coherence is too poor for self calibration to successfully determine the phase
solutions for isolated remote antennas. This ultimately causes systematic errors in the final maps; we discuss this further in Appendix B.2.
The absolute flux scale of CARMA visibility data is typically set using observations of a calibrator of 'known' flux, usually a planet. Flux
calibration using this method is expected to be accurate to roughly 20%. During our D-array observations, only a single dataset contained
observations of a primary flux calibrator (MWC3498). As described in Luszcz-Cook and de Pater (2013) we flux-calibrated this dataset
in the standard way using an assumed value of 1.86 Jy for MWC349 at 230 GHz. Then, we binned the Neptune visibility data for each
D-array dataset into 8 (u,v) bins between 0 and 80 kλ, and adjusted the flux calibration to align the binned visibilities between tracks as
much as possible. Finally, during the analysis of Neptune's disk-integrated spectrum, we determined an overall correction factor to the
8 http://cedarflat.mmarray.org/fluxcal/primary_sp_index.
htm
14
S. H. Luszcz-Cook, I. de Pater, M. Wright
flux calibration that optimizes the match between the data and our best-fit model. As discussed in Luszcz-Cook and de Pater (2013), the
continuum slope of our best-fit model is not a perfect match to the data, resulting in model fluxes that are systematically too high by ∼ 2%
near 225 GHz and too low by ∼ 2% near 236 GHz as compared to the disk-integrated measurements. For the present analysis, we scale
the D-array visibilities by the flux correction factor determined from the fit to our nominal model (Section 3). We then bin the Neptune
visibility data for each individual B-array dataset in intervals of 10 kλ at (u,v) distances for which both B- and D-array data are present
(40 -- 80 kλ). Finally, we adjust the overall flux scale for each B-array track to align the binned visibility data with the D-array data. The
real parts of the calibrated visibilities from the B and D configurations are plotted as a function of projected baseline length in Fig. 8,
where each point is the average over one observing cycle. For reference, the shape of the visibility function corresponding to a circularly
symmetric, uniform disk is a Bessel function. After flux calibration, Neptune's measured disk-integrated flux density is 12.3 Jy at 225 GHz
and 12.2 Jy at 229.5 GHz.
Fig. 8. -- The real part of the visibilities, shown as a function of projected baseline length. Each data point is the average over a single
observing cycle (8 minutes in the B configuration, 15 minutes in the D configuration). The purple points are the short baseline, D-array
data, which sample large angular scales. The B-array data (blue/green/orange for data taken in the winter of 2008/2009 -- 2010/2011,
respectively) measure smaller-scale structure. Both arrays have some baselines with projected lengths of 40-80 kλ; the D-array data at
these baseline lengths are used to flux-calibrate the B-array data as described in Appendix A.
IMAGING AND DECONVOLUTION
We produce maps of Neptune's flux in four steps: first we convert our visibility data into a set of maps of the sky intensity distribution.
Then, we estimate the noise in the maps due to atmospheric and instrumental effects. Using this estimate, we perform deconvolution to
remove the response to source structure in the side lobes of the synthesized beam. Finally, we average images to increase signal-to-noise,
and bin the data according to latitude to look for trends in Neptune's observed brightness.
Imaging
We use the Common Astronomy Software Applications (CASA) data reduction package9 to convert our visibility data into maps of
the sky intensity distribution. The frequency resolution of these maps is selected to be lower than the original channel spacing of the
visibility data in order to improve the signal-to-noise in the images: from the 62 MHz band centered on the CO line, we define 9 frequency
intervals of width 4.6 MHz. From the 500 MHz bands, we select eight 125 MHz frequency windows for mapping. The original channels are
combined in the imaging step using multi-frequency synthesis mode. As mentioned in Section 2, the imaging windows are chosen to cover
the frequencies where we have the most data. The frequencies of these imaging windows are listed in Table 4.
We apply intermediate weighting to the visibilities with a Briggs visibility weighting robustness parameter (Briggs 1995) of 0.0. We use
the mosaic option to properly handle the three different primary beam patterns of the heterogeneous CARMA array. The pixel size used
in the maps is 0.09"; the shape of the synthesized beam varies between images, with a FWHM in the range of 0.33 -- 0.37" x 0.37 -- 0.39".
Calibration errors and atmospheric fluctuations introduce complex-valued multiplicative errors into the data. As noted in Section 2, the
rms path errors in our data are as high as 350 µm, which is roughly 1/3 of the observing wavelength and well above CARMA's 150 µm
Error determination
9 http://casa.nrao.edu
15
criterion for good 1-mm observing conditions. We are able to correct for these errors on some baselines using self calibration (Appendix A);
however the phase coherence is too poor for this technique to be successful on isolated remote antennas, which results in increased errors
in our final maps above the value expected from thermal noise from the atmosphere and receiver system.
We can estimate the noise by subtracting a model from the data and making an image. The average rms in the 4.6 MHz model-subtracted
images is 28 mJy/beam, 2.7 times the value expected from thermal noise. The rms in the 125 MHz images is on average 7.8 mJy/beam,
3.3 times the value expected from thermal noise.
Another estimate of the noise can be be obtained by making an imaginary image from the non-Hermitian part of the data. For a real
valued sky brightness intensity, the visibility function is Hermitian. An image of the non-Hermitian part of the data is a direct measure of
the errors in the visibility data. The imaginary images show Gaussian distributions of pixel values with rms values that are in agreement
with the noise estimate found by the previous method. The noise estimates from these two methods are presented for the individual maps
in Table 4.
Imaging frequencies, listed in order of increasing distance from line center. The imaging frequency resolution is coarser than the original channel
spacing of the data. The image noise is estimated in two ways: method 1 involves measuring the rms in an image made from model-subtracted
data. Method 2 is to find the rms in an imaginary image made from the data.
TABLE 4
Center frequency Frequency offset
from line center
0.00 MHz
−4.62 MHz
4.62 MHz
-9.23 MHz
9.23 MHz
−13.8 MHz
13.9 MHz
−18.4 MHz
18.5 MHz
66.0 MHz
213 MHz
−0.937 GHz
−1.06 GHz
−4.94 GHz
−5.06 GHz
−6.09 GHz
−6.21 GHz
230.538 GHz
230.533 GHz
230.543 GHz
230.529 GHz
230.547 GHz
230.524 GHz
230.552 GHz
230.520 GHz
230.556 GHz
230.604 GHz
230.751 GHz
229.601 GHz
229.476 GHz
225.600 GHz
225.475 GHz
224.450 GHz
224.325 GHz
Image frequency
width
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
4.6 MHz
125 MHz
125 MHz
125 MHz
125 MHz
125 MHz
125 MHz
125 MHz
125 MHz
RMS, method 1
(mJy/beam)
28.2
28.6
27.8
27.9
28.7
26.7
29.1
27.7
26.9
8.13
7.64
8.39
8.57
7.58
7.21
7.41
7.44
RMS, method 2
(mJy/beam)
28.3
28.5
27.8
28.2
28.8
26.6
29.0
27.5
26.7
8.01
7.59
8.43
8.53
7.59
7.21
7.44
7.37
Deconvolution
The imaging step is followed by deconvolution, which attempts to reconstruct the true sky brightness distribution from the observed
image using the synthesized beam. Two different deconvolution algorithms are commonly used: an iterative point source subtraction
algorithm, CLEAN which is well-matched for deconvolving compact source structures, and maximum entropy, a gradient search algorithm,
which maximizes the fit to an a-priori image, in a least squares fit to the (u,v) data. After investigating both deconvolution strategies
and several variations of CLEAN (Appendix C), we deconvolve our synthesized maps using the Clark CLEAN algorithm (Clark 1980) with a
gain factor of 0.05, cleaning down to the rms noise level of the image. A starting CLEAN model is provided based on the radiative transfer
solution for the nominal CO profile from Luszcz-Cook and de Pater (2013) (Section 3). We note that the resulting maps do not appear to
be affected by the details of the starting CLEAN model. For a more complete comparison of the imaging and deconvolution strategies tried,
see Appendix C.
Averaging
After the initial mapping procedure, we average deconvolved images where possible to further improve the signal-to-noise: the contribution
function (Fig. 1) shows that at an offset of 5 GHz from line center, the maps are almost entirely due to continuum emission from pressures
of 1-5 bar. We therefore average the four maps at offsets of 5-6 GHz from line center to produce a single continuum image (Fig. 3). We
also combine the maps at 229.48 and 229.60 GHz (1.04 and 0.96 GHz from line center, respectively), giving us three high signal-to-noise
images at offsets of 66 MHz, 213 MHz and 1 GHz from line center (Fig. 4). In the core of the CO line where we have an imaging resolution
of 4.6 MHz, we average maps with the same absolute frequency offset from line center (Fig. 5). The contribution functions for these image
pairs are very similar, so this allows us to increase our signal to noise while maintaining sensitivity to emission from different atmospheric
pressures.
The noise in each of these averaged maps is approximated as the root mean square sum of the errors in the original maps, divided by
the square root of the number of input images. The higher of the two noise estimates for each input image (Table 4) is used. We note that
averaging of images takes place after cleaning, which is a non-linear process. We also caution that systematic errors, for example those
caused by gain calibration errors, will not decrease with more averaging. This can cause our errors to underestimate the true noise in the
images. Therefore, the noise values derived for the final, averaged maps should be taken as best estimates. These values are presented in
Table 2.
Longitudinal brightness variations due to changes in the atmospheric properties should be averaged out in the long data integrations,
but latitudinal variations may reflect meridional trends in the composition and/or temperature. To look for meaningful trends and further
increase signal to noise, we bin the data in each map in bins of latitude. We use the ephemeris data from JPL Horizons10 to get the value
of the latitude that each pixel location in the image would have if it were not convolved with the synthesized beam. To determine the
variation of the intensity in the maps with physical location on the planet, we select 7 non-overlapping latitude bins; the width of each bin
is approximately equal to one beam, such that data separated by at least one bin will be largely uncorrelated. For each latitude bin, we
make a model mask: a pixel in the mask is set to 1 if the latitude at that pixel location is within the latitude bin range; otherwise it is set
to 0. We then convolve the mask with the synthesized beam, and multiply this mask by the data. The total of this masked image divided
by the number of pixels in the mask gives a weighted average of the data over the desired latitude range. The error for each binned point
10 http://ssd.jpl.nasa.gov/?horizons
16
S. H. Luszcz-Cook, I. de Pater, M. Wright
is estimated as the image noise in mJy/beam divided by the square root of the number of beams within the binned region, which varies
from just over 1 beam for the bins centered at 80◦S and 40◦N, to nearly 6 beams in the bin centered at 20◦S. For direct comparison with
the models, we convolve the models with the CLEAN beam and multiply by the same masks.
COMPARISON OF DECONVOLUTION TECHNIQUES
Neptune, with its smooth, bright disk and sharply defined edges, presents a challenge for imaging and deconvolution. In order to evaluate
the effect and significance of our weighting function and deconvolution technique, we perform a series of comparisons using a subset of
our data that spans 437.5 MHz in frequency, centered at 225.74 GHz (nearly 5 GHz from the center of the CO (2-1) line). These data
represent an instance where there is overlap between the D-array data and all epochs of B-array data. To determine the importance of the
short-spacing (D-array) data in the final maps, we imaged the B-array data separately, in addition to imaging the full combined data set.
The synthesized maps from both data subsets are shown in Figs. 9 and 10, with three different weighting functions (natural, uniform and
intermediate) applied. Natural weighting (top), which weights each visibility by the inverse of its noise variance, gives the best sensitivity,
at the expense of the shape of the synthesized beam and sidelobe levels. Uniform weighting (middle) adjusts the weight of each visibility
so that the density of visibilities is uniform across the (u,v) plane. This minimizes sidelobe levels, but increases the noise level in the map.
Robust weighting with a Briggs visibility weighting robustness parameter (Briggs 1995) of 0.0 (bottom) is a good compromise between the
two; maintaining a well-behaved beam shape with less increase to the noise level.
Fig. 9. -- Synthesized images, B-array (long baseline) data only. Made using natural weighting (top), uniform weighting (middle) and
robust weighting with a Briggs visibility weighting robustness parameter (Briggs 1995) of 0.0 (bottom). On the left we show the full maps;
on the right we zoom in on the planet. The beam is indicated by the red oval in the bottom left corner of each image.
We experimented with several deconvolution techniques; six of them are presented here. The results are summarized in Table 5 and
shown in Figs. 11 - 14.
f
(a) Classical CLEAN no clean region: We used the CASA implementation of the Clark CLEAN routine with a gain factor of 0.05. A maximum
of 100,000 iterations were permitted, unless a clean threshold of 8.6 mJy/beam was reached. This threshold was selected to be twice
the measured rms for the synthesized map. We find that a significant amount of noise is cleaned, particularly in the case of the B-array
data subset, for which a strong negative "clean bowl" is present due to the missing short spacing information. As a result, the residual
map is artificially low, and only 3/4 of the total flux from the planet is recovered when only B-array data are included.
(b) Classical CLEAN, inside a circle of radius 1.5": We restrict CLEAN to a circular region of radius 1.5", just outside of the planet. Since
most of the flux within this region is source (not noise), we use a lower clean threshold of 4.3 mJy, equal to the measured rms of the
map. This technique recovers more than 95% of the expected (model) flux from Neptune, even when the D-array data are not included.
However, the noise outside of the CLEAN region is higher than in many of the other test methods, and the disk appears very 'lumpy'
in the final CLEAN map, presumably because the flux is being reproduced by a set of point sources, which is unrealistic in the case of
Neptune.
17
Fig. 10. -- Synthesized images, using all data (B and D arrays). Made using natural weighting (top), uniform weighting (middle) and
robust weighting with a Briggs visibility weighting robustness parameter (Briggs 1995) of 0.0 (bottom). On the left we show the full maps;
on the right we zoom in on the planet. The beam is indicated by the red oval in the bottom left corner of each image.
TABLE 5
Comparison of deconvolution strategies.
Algorithm
Region
Input model Threshold
(a) CLEAN
-
(b) CLEAN
1.5" circle
squared
(c) CLEAN
squared
(d) CLEAN
(e) MSCLEAN
2.5" circle
(f) Max. entropy 1.5" circle
-
-
nominal
flat
-
-
(mJy/beam)
8.6
4.3
4.3
4.3
4.3
4.3
B-array data only
Fluxa RMSb
(Jy)
9.10
11.56
12.62
10.88
13.34
12.44
(mJy/beam)
2.1
2.0
1.5
1.4
1.9
5.5
Full dataset
NITERc Fluxa RMS
NITERc
6810
8451
6477
11344
680
88
(Jy)
12.26
12.25
12.24
12.24
12.24
12.50
(mJy/beam)
2.0
2.0
1.4
1.6
2.0
6.2
2180
2579
9918
8168
628
71
aTotal flux recovered; nominal clean model gives a total flux of 12.740 Jy
bstandard deviation of the full residual map
cnumber of CLEAN or maximum entropy iterations performed
dsquare extends just beyond the Neptune model disk- approximately 2.3" on a side
(c) Classical CLEAN, starting with best-guess clean model: CLEAN is given a starting model, which is taken from our radiative transfer model
described in Section 3. The input model uses 0.005" pixels that cover a square region that extends just beyond Neptune's limb. CLEAN
then proceeds in the same way as before, adding additional point source components (both positive and negative) within the spatial
range the starting model until the CLEAN threshold is reached. In general, we find this approach gives a good result: the noise is low
across the map (even beyond the CLEAN region), and the disk appears far smoother than for the previous two techniques.
(d) Classical CLEAN, starting with a flat clean model: To address the possibility that our choice of input model has an unexpected effect on
the output map (perhaps forcing the data to match our nominal model), we repeat case (c) using a flat input model, which is a disk of
constant intensity, equal to the lowest intensity value of the best guess model in (c). We find that without including the D-array data,
some of the negative bowl from the missing short-spacing data remains with this technique, so that the measured flux from Neptune
is low. However, when the D-array data are included, a flat input starting model produces a very similar map to the more detailed
model in (c), and the noise properties of the map are similar for the two starting model cases.
(e) Multiscale CLEAN: As an alternative to classical CLEAN we try the CASA implementation of multi-scale CLEAN. Rather than modeling
the sky brightness with a set of point-sources, multi-scale CLEAN models the sky with components of several different size scales. For a
18
S. H. Luszcz-Cook, I. de Pater, M. Wright
Fig. 11. -- Comparison of deconvolution techniques, for the B-array data only. Discussion of techniques (a)-(f) is given in Appendix
C. For each technique, the deconvolved map (top) residual map (middle) and deconvolution model (bottom) are shown. The image size
is the same as in Fig. 10, right. The deconvolution model is the model of the sky that includes any input (starting) model component
plus the components determined by the deconvolution. The deconvolved map is the synthesized map, less the model convolved by the
synthesized beam, plus the model convolved by a gaussian fit to the synthesized beam. The residual map is the synthesized map, less the
model convolved by the synthesized beam.
Fig. 12. -- Comparison of maps from deconvolution techniques (a)-(f), for the B-array data only. Plots are made by slicing horizontally
through the center of the maps, and shown as relative pixel numbers, where the center pixel is at 0 (each pixel has a width of 0.09"). Also
shown is our nominal model (black) for the data. In the bottom plot, we subtract the nominal model from each of the map slices.
detailed comparison of the multi-scale CLEAN and traditional CLEAN algorithms, see Rich et al. (2008). We use a gain of 0.3; the threshold
is set to 4.3 mJy. Five size scales are specified, ranging from a point source up to the diameter of Neptune. Since multi-scale CLEAN
will only place model components such that they are entirely contained within the cleaning region, we specify a larger cleaning radius
of 2.5". We find that multi-scale CLEAN is able to recover Neptune's flux even when the short spacing data were omitted. Multi-scale
CLEAN also requires significantly fewer iterations to converge than classical CLEAN , and agrees well with (c) without depending at all on
an input CLEAN model. However, the noise level outside of the cleaning region is somewhat higher than in (c) and (d).
(f) Maximum Entropy: We use the mosmem routine in MIRIAD, which is an implementation of the Maximum Entropy Method (MEM).
This algorithm, which is an alternative to CLEAN produces a smooth positive image which maximizes the fit to an a-priori image, in
a least squares fit to the (u,v) data. We restrict the deconvolution to a circle of radius 1.5", and specify the total flux based on the
expected value from the model. We allow for as many as 5000 iterations to reduce the image residuals to 4.3 mJy/beam. We found
that mosmem converged quickly; however, this approach had the highest residual level of any of the methods we tried.
From this comparison, we conclude that when the D-array (short baseline) data are included, there is no significant difference in the final
maps from the different deconvolution methods. In particular, while a starting input CLEAN model appears to improve the appearance of
the final maps, it does not artificially force the resulting maps to agree with the starting model, as long as the short baseline data are
included. Table 5 lists the total flux densities in the maps. The maps produced using the full dataset produce a final flux density that is
consistently lower than the model predicts. This result is consistent with the disk-integrated spectrum described in Luszcz-Cook and de
0200400600 mJy/beammodel(a)(b)(c)(d)(e)(f)−40−2002040relative pixel #−100−5005019
Fig. 13. -- Same as Fig. 11, except for all data (B and D arrays).
Fig. 14. -- Same as Fig. 12, except for all data (B and D arrays).
Pater (2013), which indicates that at 225 GHz the model over-predicts Neptune's total flux. When we image only the B-array data, the
use of a starting model is more influential, reducing the CLEAN bowl and increasing the total flux.
As discussed in Appendix B.2, calibration errors and atmospheric fluctuations introduce amplitude and phase errors into the visibility
data. The 'ringing' in the final maps is not an artifact of deconvolution, but is a result of these errors. This is supported by the fact that
the rippling is observed in maps produced using both CLEAN and Maximum Entropy deconvolution methods, and is in fact observable in the
synthesized images (prior to deconvolution). These ripples have an amplitude as high as 1 mJy/pixel, and the average deviation between
map (c) and a smooth model (scaled to the data amplitude) is 5.7 mJy/beam.
Since the maps are not overly influenced by the inclusion of an input model, the appearance of map (c) is best, and the residuals in case
(c) are the lowest, we select method (c) for our image deconvolution.
0200400600 mJy/beammodel(a)(b)(c)(d)(e)(f)−40−2002040relative pixel #−100−50050 |
1905.06035 | 1 | 1905 | 2019-05-15T08:59:12 | Ground-based follow-up observations of TRAPPIST-1 transits in the near-infrared | [
"astro-ph.EP",
"astro-ph.SR"
] | The TRAPPIST-1 planetary system is a favorable target for the atmospheric characterization of temperate earth-sized exoplanets by means of transmission spectroscopy with the forthcoming James Webb Space Telescope (JWST). A possible obstacle to this technique could come from the photospheric heterogeneity of the host star that could affect planetary signatures in the transit transmission spectra. To constrain further this possibility, we gathered an extensive photometric data set of 25 TRAPPIST-1 transits observed in the near-IR J band (1.2 $\mu$m) with the UKIRT and the AAT, and in the NB2090 band (2.1 $\mu$m) with the VLT during the period 2015-2018. In our analysis of these data, we used a special strategy aiming to ensure uniformity in our measurements and robustness in our conclusions. We reach a photometric precision of $\sim0.003$ (RMS of the residuals), and we detect no significant temporal variations of transit depths of TRAPPIST-1 b, c, e, and g over the period of three years. The few transit depths measured for planets d and f hint towards some level of variability, but more measurements will be required for confirmation. Our depth measurements for planets b and c disagree with the stellar contamination spectra originating from the possible existence of bright spots of temperature 4500 K. We report updated transmission spectra for the six inner planets of the system which are globally flat for planets b and g and some structures are seen for planets c, d, e, and f. | astro-ph.EP | astro-ph | MNRAS 000, 1 -- 21 (2019)
Preprint 16 May 2019
Compiled using MNRAS LATEX style file v3.0
Ground-based follow-up observations of TRAPPIST-1
transits in the near-infrared
A. Y. Burdanov,1(cid:63) S. M. Lederer2, M. Gillon1, L. Delrez3, E. Ducrot1, J. de Wit4,
E. Jehin5, A. H. M. J. Triaud6, C. Lidman7, L. Spitler8,9, B.-O. Demory10, D. Queloz3
and V. Van Grootel5
1Astrobiology Research Unit, Universit´e de Li`ege, All´ee du 6 Aout 19C, 4000 Li`ege, Belgium
2NASA Johnson Space Center, 2101 NASA Parkway, Houston, TX 77058, USA
3Cavendish Laboratory, JJ Thomson Avenue, Cambridge, CB3 0H3, UK
4Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Avenue,
Cambridge, MA 02139, USA
5Space Sciences, Technologies and Astrophysics Research (STAR) Institute, Universit´e de Li`ege, All´ee du 6 Aout 19C, 4000 Li`ege, Belgium
6School of Physics & Astronomy, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
7The Research School of Astronomy and Astrophysics, Australian National University, ACT 2601, Australia
8Research Centre for Astronomy, Astrophysics & Astrophotonics, Macquarie University, Sydney, NSW 2109, Australia
9Department of Physics & Astronomy, Macquarie University, Sydney, NSW 2109, Australia
10University of Bern, Center for Space and Habitability, Sidlerstrasse 5, CH-3012 Bern, Switzerland
Accepted 2019 May 14. Received 2019 May 9; in original form 2019 April 3
ABSTRACT
The TRAPPIST-1 planetary system is a favorable target for the atmospheric charac-
terization of temperate earth-sized exoplanets by means of transmission spectroscopy
with the forthcoming James Webb Space Telescope (JWST). A possible obstacle to this
technique could come from the photospheric heterogeneity of the host star that could
affect planetary signatures in the transit transmission spectra. To constrain further
this possibility, we gathered an extensive photometric data set of 25 TRAPPIST-1
transits observed in the near-IR J band (1.2 µm) with the UKIRT and the AAT,
and in the NB2090 band (2.1 µm) with the VLT during the period 2015-2018. In our
analysis of these data, we used a special strategy aiming to ensure uniformity in our
measurements and robustness in our conclusions. We reach a photometric precision of
∼ 0.003 (RMS of the residuals), and we detect no significant temporal variations of
transit depths of TRAPPIST-1 b, c, e, and g over the period of three years. The few
transit depths measured for planets d and f hint towards some level of variability, but
more measurements will be required for confirmation. Our depth measurements for
planets b and c disagree with the stellar contamination spectra originating from the
possible existence of bright spots of temperature 4500 K. We report updated transmis-
sion spectra for the six inner planets of the system which are globally flat for planets
b and g and some structures are seen for planets c, d, e, and f.
techniques: methods: data analysis -- techniques: photometric -- stars:
Key words:
individual: TRAPPIST-1 -- planets and satellites: atmospheres -- infrared: planetary
systems -- infrared: stars.
9
1
0
2
y
a
M
5
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
3
0
6
0
.
5
0
9
1
:
v
i
X
r
a
1 INTRODUCTION
TRAPPIST-1 is a ∼ 0.09M(cid:12) ultracool dwarf star 12.4 pc
away in the Aquarius constellation (Gizis et al. 2000; Van
Grootel et al. 2018). Photometric monitoring revealed that it
(cid:63) E-mail: [email protected]
© 2019 The Authors
hosts a compact resonant system composed of seven transit-
ing Earth-sized planets (Gillon et al. 2016; Gillon et al. 2017;
Luger et al. 2017). Dynamical modeling of this system based
on timing variations of its planets' transits resulted in strong
constraints on the planetary masses which, when combined
with the radii measured from transits, point toward rocky
compositions with sizable volatile contents (Grimm et al.
2
A. Burdanov et al.
2018). With stellar irradiations ranging from 0.1 to 4 times
solar, these seven planets can be qualified as "temperate",
and three of them (planets e, f, and g) orbit within the "hab-
itable zone" of the host star (Gillon et al. 2017).
Due to the combination of relatively large infrared
brightness (J ∼ 11.4, H ∼ 10.7, K ∼ 10.3) and the small
size (∼ 0.12R(cid:12)) of their host star, the TRAPPIST-1 plan-
ets are especially favorable targets for the detailed explo-
ration of their atmospheres with the future ground-based
extremely large telescopes and James Webb Space Tele-
scope (JWST) (Rodler & L´opez-Morales 2014; Barstow &
Irwin 2016; Gillon et al. 2017; Morley et al. 2017; Lincowski
et al. 2018). First reconnaissance observations with the Hub-
ble Space Telescope (HST) by de Wit et al. (2016, 2018)
showed no hint of clear H/He dominated atmospheres for the
six inner planets. Further studies reassessed these findings:
Moran et al. (2018) supported the presence of secondary
volatile-rich atmospheres for planets d, e and f using revised
masses (while not entirely ruling out a cloud-free H-rich at-
mosphere for TRAPPIST-1 e and f) and Wakeford et al.
(2019) ruled out a clear solar H2/He dominated atmosphere
for TRAPPIST-1 g.
A possible obstacle on the way for a proper and ro-
bust characterization of the atmospheres of the TRAPPIST-
1 planets was presented by Apai et al. (2018) and Rackham
et al. (2018), who argue that the photospheric heterogeneity
of the host star, and specifically spots and faculae, could al-
ter, hide, or even mimic planetary signatures in the transit
transmission spectra. Taking into account this possible stel-
lar contamination, and basing their re-analysis on existing
HST near-IR data, Zhang et al. (2018) predicted consider-
able changes of the transit depths of TRAPPIST-1 planets
with wavelength. Work by Ducrot et al. (2018) disproved
these predictions by producing featureless broadband trans-
mission spectra for the planets in the 0.8-4.5 µm spectral
range, and showing an absence of significant temporal varia-
tions of the transit depths in the visible. Furthermore, Delrez
et al. (2018) showed that the 3.3 d periodic 1% photometric
modulation detected in the K2 optical data set (Luger et al.
2017) is not present in Spitzer 4.5 µm observations, and
that transit depth measurements do not show any hint of a
significant stellar contamination in this spectral range. The
same conclusion was reached by Morris et al. (2018b) using
a "self-contamination" approach based on the measurement
of transit egress and ingress durations from the same Spitzer
data set.
Based on the presence of a 3.3 d photometric variability
in K2 data and its absence in Spitzer mid-IR data, Mor-
ris et al. (2018a) proposed that the photospheric inhomo-
geneities of TRAPPIST-1 could be dominated by a few fac-
ulae (i.e. bright spots) with characteristic temperatures in
excess of ∼ 1800 K relative to the effective temperature of
the star (∼ 2500 K). This is globally consistent with the
analysis of Rackham et al. (2018) and Zhang et al. (2018)
who inferred whole-disk spot and faculae coverage fractions
of ∼10 and ∼ 50% respectively, but with a much lower tem-
perature excess of (∼ 500K) for the faculae. According to
Ducrot et al. (2018), their analyses favor two scenarios of
TRAPPIST-1 photosphere: with a prevalence of a several
high-latitude cold and large spots, or alternatively, a few
hot and small spots.
Globally, all these studies do not yet present a consis-
tent picture of the photospheric properties of TRAPPSIT-1,
even at the observational level. For instance, the 1.1-1.7 µm
combined transmission spectra presented by Ducrot et al.
(2018) do not show any significant features, while the spec-
tra obtained by Zhang et al. (2018) from the very same
HST data set shows a drop around 1.4 µm attributed by
the authors to an inverted water absorption feature caused
by stellar contamination. Both studies agree that the HST
transit depths are globally deeper than those measured at
other wavelengths. Nevertheless, as outlined by Ducrot et al.
(2018), the origins of these larger transit depths could be in-
strumental, as the HST/WFC3 systematic effects combined
with the low-Earth orbit of HST make it very difficult to nor-
malize transit light curves as short as those of TRAPPIST-1
planets (see Subsection 3.2 in Ducrot et al. 2018). The uncer-
tainties affecting the absolute values of the transits measured
by HST in the near-IR is unfortunate, as this spectral range
encompasses the peak of the spectral energy distribution of
TRAPPIST-1 (see Fig. 3 in Morris et al. 2018a) and could
bring strong constraints on the photospheric properties of
TRAPPIST-1 and on the impact of stellar contamination to
be expected for upcoming JWST observations of the planets.
As Earth's atmosphere is partially transparent in near-IR,
it is therefore highly desirable to observe as many transits
of TRAPPIST-1 planets as possible in this spectral range at
high precision with ground-based telescopes.
In this paper, we present an extensive photometric data
set of 25 TRAPPIST-1 transits gathered in the near-IR
J band (1.2 µm) with the 3.8-m United Kingdom Infra-Red
Telescope (UKIRT) and the 3.9-m Anglo-Australian Tele-
scope (AAT), and in the NB2090 band (2.1 µm) with the
ESO Very Large Telescope (VLT) during the period 2015-
2018. In our analysis of these data, we used a special strat-
egy (see Subsections 3.2 and 3.3) aiming to ensure unifor-
mity in our measurements and robustness in our conclusions.
This special attention is motivated by the inherent com-
plexity of ground-based near-IR data reduction appearing
as correlations of deduced transit depths with photometric
aperture sizes and comparison stars (for a comprehensive
review of this topic see Croll et al. 2015). We also report
precise timings of each transit that should further constrain
the masses of the planets via the transit-timing variations
method (Gillon et al. 2017; Grimm et al. 2018).
The rest of the paper is divided into four sections. First,
we describe our observational data set in Section 2. Section 3
is devoted to the reduction and analysis of these data. We
present and discuss our results in Section 4, and we outline
our findings in the Conclusions section.
2 INSTRUMENTS AND OBSERVATIONS
We acquired more than 30 transits of TRAPPIST-1 planets
with UKIRT, VLT, and AAT in the period from December
2015 to July 2018 through different observing programs (see
Acknowledgements section for a full list of programs). For
our analysis, we considered only 25 transits which satisfied
the following two conditions: (1) isolated transits, i.e. not
blended with the transit of another planet, and (2) taken
under relatively good observing conditions, meaning trans-
parency variations in the Earth's atmosphere were less than
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
3
2.2 VLT/HAWK-I
We used the HAWK-I cryogenic wide-field imager installed
on Unit Telescope 4 (Yepun) of the ESO VLT at Paranal
observatory to observe transits of TRAPPIST-1 b and c in
2015 and 2017. The HAWK-I imager is composed of four
2048 × 2048 Hawaii 2RG chips (Siebenmorgen et al. 2011).
Each chip provides an image scale of 0.106 arcsec pixel−1 re-
sulting in a 217× 217 arcsec2 FoV. TRAPPIST-1 was placed
in the corner of the quadrant Q3 (chip #79) to allow three
additional stars to be simultaneously imaged on the chip for
use as comparison stars. Non-Destructive Read (NDR) was
used with a 3 second exposure time and 12 sub-integrations.
All transits were observed in narrow-band filter NB2090
which has a central wavelength of 2.095 µm and width of
0.020 µm. The small width of this filter minimizes the effect
of differential extinction, while the combination of its central
wavelength and bandwidth eliminates large absorption and
emission bands present in the K band2.
2.3 AAT/IRIS2
We observed two transits of TRAPPIST-1 b with the IRIS2
IR-imager installed on the AAT 4-m telescope at the Sid-
ing Spring Observatory. IRIS2 IR-imager consists of one
1024 × 1024 Rockwell Hawaii-II detector which has a FoV
of 7.7 × 7.7 arcmin2 and a pixel scale of 0.45 arcsec pixel−1
(Tinney et al. 2004). We used 9 second exposure times and
observations were done in the J band. The telescope was
pointed in such a way as to prevent TRAPPIST-1 and com-
parison stars from falling in the upper right quadrant of
IRIS2, which had excessive noise.
3 DATA REDUCTION AND ANALYSIS
In this section, we describe the entire data handling pro-
cess which consisted of a preliminary reduction of the im-
ages, photometric extraction of the stellar fluxes, perform-
ing differential photometry, and deducing transit parame-
ters with the use of an adaptive Markov Chain Monte Carlo
(MCMC) code (Gillon et al. 2012, 2014). Each transit was
first analysed individually to obtain its parameters and to
search for possible temporal variability (subsection 3.3). For
each planet, global MCMC analyses of all the transits were
then performed (section 3.4), although separately for each
filter.
3.1 Image calibrations
3.1.1 UKIRT/WFCAM
All
images obtained with the UKIRT telescope were
pre-processed by the Cambridge Astronomy Survey Unit
(CASU) and then were downloaded from the WFCAM Sci-
ence Archive (WSA). Reduction steps completed by CASU
include: dark-correction, flat-fielding, gain-correction and
2 https://www.eso.org/sci/facilities/paranal/
instruments/hawki/inst.html
Figure 1. Distribution of the TRAPPIST-1 planet transits ana-
lyzed in this work.
30%. The distribution of these 25 transits among the differ-
ent TRAPPIST-1 planets is shown in Fig. 1.
All transits were observed in "staring mode", i.e. with-
out dithering of the telescope. However, before and af-
ter the scientific sequence, sky flat images were acquired
with dithering to construct a proper master sky flat image.
An observing log is presented in Table 1, and we outline
instrument-specific information below.
2.1 UKIRT/WFCAM
The WFCAM near-IR imager of the UKIRT 3.8-meter tele-
scope located on the summit of Mauna Kea was used to ob-
serve transits of TRAPPIST-1 b, c, d, e, f and g in the broad-
band J filter1. Those observations occurred between 2015
and 2018. The WFCAM imager consists of four 2048 × 2048
Rockwell Hawaii-II detectors with a field of view (FoV)
of 13.65 × 13.65 arcmin2 each and image scale of 0.4 arc-
sec pixel−1 (Casali et al. 2007). TRAPPIST-1 was placed in
quadrant 3 (array ID number #76) as this is the cleanest of
the four detectors.
Observations were made in Correlated Double Sampling
(CDS) mode which is the default read mode used for all
broad-band observations. These observations comprised co-
adding five x 1-second exposures throughout the undithered
scientific sequence. The sequence was extended long enough
- 1 to 2.5 hours on each side of the transit, depending upon
how well known the orbital period of the target planet's
transit was known - to ensure that a pre- and post-transit,
star-only baseline was acquired. For example, a 1 h tran-
sit of planet f, plus 1.5 h to capture the baseline out-of-
transit prior to and following the transit, resulted in a 4 h
total undithered scientific sequence. The dithered sky flat se-
quence before and after each full scientific transit sequence
was composed of capturing five x 1-second co-added expo-
sures throughout the 2.5 min sequence.
1 http://casu.ast.cam.ac.uk/surveys-projects/wfcam/
technical/filter-set
MNRAS 000, 1 -- 21 (2019)
bcdefghPlanet0246810Number of transits4
A. Burdanov et al.
Table 1. Observing log: observations are grouped by planets and sorted chronologically in each group.
ID
Date of start
of the night
Planet Telescope/Instrum.
Filter
Duration
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
2015 Nov 07
2015 Dec 05
2015 Dec 08
2016 Jul 29
2016 Aug 01
2016 Oct 17
2016 Oct 20
2017 Jul 07
2017 Nov 30
2018 Jul 06
2015 Dec 06
2016 Jul 18
2016 Jul 30
2017 Nov 21
2017 Aug 16
2017 Aug 20
2017 Oct 28
2017 Sep 11
2018 Jul 13
2017 Jul 31
2017 Sep 15
2017 Oct 22
2017 Sep 02
2017 Dec 10
2018 Jun 13
b
b
b
b
b
b
b
b
b
b
c
c
c
c
d
d
d
e
e
f
f
f
g
g
g
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
AAT/IRIS2
AAT/IRIS2
UKIRT/WFCAM
J
J
J
J
J
J
J
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
J
J
J
J
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
J
J
J
J
J
J
J
J
J
J
J
∼ 4.0 h
∼ 3.5 h
∼ 3.6 h
∼ 5.6 h
∼ 3.2 h
∼ 3.7 h
∼ 3.4 h
∼ 2.0 h
∼ 2.8 h
∼ 1.8 h
∼ 3.8 h
∼ 4.9 h
∼ 3.3 h
∼ 3.1 h
∼ 3.0 h
∼ 3.0 h
∼ 2.9 h
∼ 2.4 h
∼ 2.9 h
∼ 3.8 h
∼ 3.2 h
∼ 3.9 h
∼ 5.0 h
∼ 3.5 h
∼ 2.6 h
Exposure time
(N×EXP)b
12×3.0 s
3×2.0 s
5×1.0 s
5×1.0 s
5×1.0 s
1×9.0 s
1×9.0 s
5×1.0 s
12×3.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
12×3.1 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
5×1.0 s
Sky transparencya
Remarks
Clear
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Photometric
Thin cirrus
Photometric
Thin cirrus
Photometric
Photometric
Photometric
-
-
Partial
-
-
9 min gap
during bottom
Airmass >2
from 7682.13
Short OOTc ,
gap during egress
-
Short OOT, gap
before engress
-
-
-
-
Short gaps during
egress and bottom
Short gap during
bottom
-
Short gap
after egress
-
-
Short gap
before ingress
-
-
-
Short OOT, gap
during ingress
a According to the ESO definitions (https://www.eso.org/sci/observing/phase2/ObsConditions.html)
b N is the number of exposures and EXP is the individual exposure times
c out-of-transit observations
decurtaining (a specific correction for WFCAM3 data). No
linearity correction was applied as the system is linear to
< 1% up to the saturation regime4 (∼ 40 000 counts/pixel)
and maximum counts on TRAPPIST-1, which was the
brightest star in the FoV, never exceeded 15 000 counts.
Then the values of the bad pixels (deviating by 3σ when
classified as background sky values and 40σ for stars, where
σ is sky background noise) were replaced with the median
values of the neighbouring pixels.
3 http://casu.ast.cam.ac.uk/surveys-projects/wfcam/
technical/decurtaining
4 http://casu.ast.cam.ac.uk/surveys-projects/wfcam/
technical/linearity
3.1.2 VLT/HAWK-I
Raw scientific images and corresponding processed calibra-
tion images were downloaded from the ESO archive. Be-
fore preliminary data reduction, sub-images from quadrant
Q3 were extracted, and all subsequent steps were performed
only for quadrant Q3. We used only one quadrant as using
reference stars located on other chips reduces photometric
precision (Lendl et al. 2013). This also minimized compu-
tational time for data reduction. Processed calibration files
from the ESO archive were used for dark and flat field correc-
tions with the PyRAF/ccdproc5 module. Then we followed
5 http://stsdas.stsci.edu/cgi-bin/gethelp.cgi?ccdproc
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
5
the same procedure outlined for the UKIRT data to deal
with the bad pixels. As with UKIRT data, no linearity cor-
rection was applied because the detector is linear to < 1%
below 30 000 counts and maximum counts on TRAPPIST-1
were below 20 000.
3.1.3 AAT/IRIS2
All the images from AAT were treated similarly to the data
from VLT except for the manual creation of the master dark
and flat images.
3.2 Photometric extraction
After the preliminary reduction steps described above, we
applied the subsequent procedures to the data from all the
telescopes.
All images of each observing run were aligned with
respect to the first image of the run (typical X and Y
shifts were less than 1-2 pixels). Then we created a me-
dian stacked image from all the aligned images to run a
star identification algorithm. Positions of the stars on the
stacked image were identified with intensity-weighted cen-
troids. Then we extracted fluxes of the stars on each image
with eight different circular apertures, sky buffers, and sky
annuli using DAOPHOT (Tody 1986). Aperture sizes were
defined as 1, 1.5, 2, 2.5, 3, 4, 5, 6× FWHM, where FWHM is the
mean full width at half maximum of the star's point spread
function (PSF) in the image. For each star, the sky back-
ground was measured in an annulus beyond the stellar aper-
ture using the median sky fitting algorithm implemented in
PyRAF/DAOPHOT. Radius of the annulus was defined as
3× FWHM, and its width as 5× FWHM.
At that stage, differential photometry was carried out
using a custom-made code: for all apertures, a given target
star T and a list of comparison stars (C1, C2,, etc...), we
computed the ratio FT/(FC1 +FC2 +...) for each image, where
F is the flux of a star corrected for the sky background. We
derived photometric uncertainties with the use of a "CCD
equation" (Howell 2006) taking into account dark current,
read-out noise, stellar scintillation and sky Poisson noise.
At first, we followed the approach to differential pho-
tometry and transit light curve analysis that is generally
used for ground-based transit photometry obtained in the
visible (Ducrot et al. 2018). The best aperture size and com-
bination of comparison stars were selected on the basis of
the minimization of the out-of-transit (OOT) scatter of the
light curve. Similar to other scientists dealing with near-IR
ground-based observations, we found out that this approach
is not optimal as it appears to induce correlations of tran-
sit depths deduced from the MCMC analysis with aperture
size and with the used comparison stars (see Croll et al.
2015 and Clark et al. 2018). While the exact sources of these
correlations are unknown, we suspect that more likely they
are coming from a combination of these factors: brightness
and color differences of the TRAPPIST-1 and reference stars
(the target star is the brightest and the reddest star in the
FoV while reference stars are fainter and much bluer); spa-
tial separation of the target and reference stars on the sky
and on the detector, which make effects of the detector sys-
tematics and of the Earth's atmosphere display themselves
differently for the target and reference stars.
MNRAS 000, 1 -- 21 (2019)
We found that different combinations of the reference
stars and aperture sizes can be equally good in terms of the
OOT scatter, but give significantly different transit depths
for a given observing run. Thus, if such an approach is ap-
plied to the entire TRAPPIST-1 data set, then one can not
truly discriminate temporal variations of the transit depths
caused by this near-IR photometry effect, or caused by phys-
ical stellar contamination coming from the star, e.g., from
unocculted star spots, or caused by some other astrophysi-
cal process. Therefore, selection of the photometric aperture
size and suitable comparison stars is a non-trivial and impor-
tant step. In the next section, we describe how we rigorously
approached our analysis to yield credible and robust tran-
sit depths, as well as transit timings, durations and impact
parameters.
3.3 Individual analysis
One way to robustly determine output metrics is to begin
by producing a large set of differential light curves for each
transit, based on different combinations of apertures and
comparison stars. Each light curve within this set is then
analyzed with the MCMC code and the metric RMS× β2
r
(proposed by Croll et al. 2015) calculated for each. Only
those combinations whose output metric meet the criteria
of RMS× β2
the minimum metric) are selected.
r]min (less than 15% above
r < 1.15 × [RMS × β2
Here RMS is the root mean square of the photometric
residuals after the subtraction of the transit best-fit model
and βr is a quantitative factor used to assess the amount of
red (correlated) noise in our time-series (Winn et al. 2008).
After this selection is done, posterior probability distribu-
tion functions (PDFs) from the MCMC runs of the selected
light curves are combined for each transit parameter. Use
of this metric allows us to find a proper balance between
small aperture radii, which tend to produce the smallest
RMS, and large aperture radii, which minimizes the time-
correlated noise (βr ∼ 1).
For our MCMC analyses, we used the quadratic limb-
darkening (LD) law coefficients u1 and u2 with a normal
prior distributions. LD coefficients values and corresponding
uncertainties were interpolated from the paper by Claret
et al. (2012). Respective values for J band were u1 = 0.19 ±
0.04 and u2 = 0.50 ± 0.12, and for the NB2090 band -- u1 =
0.23 ± 0.05 and u2 = 0.33 ± 0.08.
Each transiting light curve was modeled using the model
by Mandel & Agol (2002) multiplied by a baseline polyno-
mial model aiming to account for the correlation of the mea-
sured fluxes with the variations of external parameters such
as the x- and y-drift of the stars on a chip due to imperfect
telescope tracking, FWHM, time, airmass, sky background
values, etc. For each observing run, we first selected a base-
line model by running a relatively short chain of MCMC with
10 000 steps with different combinations of external parame-
ters on a limited set of light curves which were, in their turn,
obtained with varying combinations of comparison stars and
aperture sizes. A model giving the minimum Bayesian Infor-
mation Criterion (BIC, Schwarz 1978) was selected, and this
model worked efficiently for all the light curves within one
observing night. Typically, accounting for the drift of the
positions of the stars, for sky background variations, and
FWHM changes decreased the BIC significantly. Selected
6
A. Burdanov et al.
Table 2. Details of TRAPPIST-1 transit light curve analyses. Baseline function is a polynomial of time (t), mean FWHM of the star's
PSF in the image ( f whm), position of the star on the CCD (xy), airmass (A) and/or sky background (sky). The epoch is calculated
using the transit ephemeris reported in Delrez et al. (2018)
ID Planet Telescope/Instrum.
Filter
Number of points
Epoch
Baseline
.
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
b
b
b
b
b
b
b
b
b
b
c
c
c
c
d
d
d
e
e
f
f
f
g
g
g
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
AAT/IRIS2
AAT/IRIS2
UKIRT/WFCAM
J
J
J
J
J
J
J
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
J
J
J
J
VLT/HAWK-I
NB2090
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
UKIRT/WFCAM
J
J
J
J
J
J
J
J
J
J
J
195
880
1107
1558
898
1211
1916
584
152
516
1148
1451
1003
169
832
866
886
697
832
1123
822
950
1468
1057
715
8
26
28
183
185
236
238
410
507
651
33
126
131
329
77
78
95
57
107
32
37
41
27
35
50
t2 + f whm2 + xy2
t3 + sky2 + xy1
t1 + f whm1 + sky3 + xy1
t2 + f whm1 + xy2
t2 + f whm3 + xy1
t2 + f whm3 + xy3
t2 + f whm3 + sky1 + xy3
t2 + f whm4 + sky1 + xy1
t3 + f whm2 + sky1 + xy2
t1 + f whm1 + xy1 + sky2
t2 + f whm2 + xy1 + sky1
t2 + f whm2 + sky2 + xy2
t4 + f whm2 + xy1
t2 + f whm2 + sky2 + xy4
t2 + f whm2 + xy2
t2 + A2 + f whm1 + xy1 + sky1
t2 + f whm2 + xy1 + sky2
t2 + f whm2 + xy2 + sky1
t2 + f whm3 + xy2
t3 + f whm3 + xy2
f whm1 + xy2
t2 + f whm2 + xy2 + sky2
t2 + f whm2 + xy2
t2 + f whm4 + xy2
A2 + f whm3 + xy2
baseline functions for each observing run are presented in
Table 2.
For each light curve of every observing run, we allowed
the following parameters to vary (jump parameters): the
mid-transit time T0; the ratio of the planet's and star's areas
(Rp/R(cid:63))2, where the planetary radius is Rp and the stellar
radius is R(cid:63); the transit width W (duration from the first
to last contact); the impact parameter b(cid:48) = a cos ip/R(cid:63) as-
suming circular orbit, where a is the semi-major axis and
ip is the orbital inclination; the combinations c1 = 2u1 + u2
and c2 = u1 − 2u2 where u1 and u2 are the quadratic limb-
darkening coefficients. The orbital period for each planet was
kept fixed to the values presented in Delrez et al. (2018).
Uniform non-informative prior distributions were assumed
for all jump parameters. TRAPPIST-1's effective tempera-
ture Teff = 2516 ± 41 K, its mass M(cid:63) = 0.089 ± 0.006 M(cid:12) and
its metallicity [Fe/H]= 0.04 ± 0.08 were applied from Van
Grootel et al. (2018) with normal prior distributions.
For every transit observation, we formed differential
light curves using all possible combinations of the compari-
son stars and aperture sizes. Typically, there were 5-6 suit-
able comparison stars, and with eight apertures, it produced
400-500 individual light curves for a given transit. However,
in the case of VLT data, we had just two suitable reference
stars. For each light curve, we ran MCMC with 40 000 steps
to estimate a best-fit value for the metric RMS× β2
r , to se-
lect light curves and to compute correction factors CF to
rescale photometric error bars (see Gillon et al. 2012 for the
details). Then for each best light curve, an MCMC model
with two chains of 100 000 steps were executed with cor-
rection factors. We controlled the convergence of the chains
by applying the statistical test of Gelman & Rubin (1992)
and checking that it was less than 1.11 for each jump pa-
rameter. Posterior PDFs derived from the analysis of these
selected light curves were combined afterward for each tran-
sit parameter. Our MCMC simulations were parallelized and
were done with the use of the Consortium des ´Equipements
de Calcul Intensif (C´ECI) computing center.
Transit depths, timings, durations and impact param-
eters deduced from our analysis are shown in Table 3 and
transit depths are discussed in the Results and Discussion
section. We plot individual transit depths in Fig. 2.
By using the approach of combining PDFs of the light
curves selected basing on the RMS× β2
r metric for every
transit, our errors on transit depths tend to increase com-
pared to the analysis of a single 'best' light curve. But this
approach gave robust estimates of the transit depths, which
is critical in the assessment of their temporal variability and
associated possible stellar contamination.
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
7
Table 3. Results of the individual analyses: median values of the posterior PDFs and their respective 1-σ limits derived for the timing
T0 , depth dF, duration W and impact parameter b for each transit.
ID
Planet
Epoch
Filter
T0
(BJDTDB -- 2450000)
σT0
dF
σd F
(per cent)
W
(min)
σW
b
σb
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
b
b
b
b
b
b
b
b
b
b
c
c
c
c
d
d
d
e
e
f
f
f
g
g
g
8
26
28
183
185
236
238
410
507
651
33
126
131
329
77
78
95
57
107
32
37
41
27
35
50
NB2090
J
J
J
J
J
J
J
NB2090
J
J
J
J
NB2090
J
J
J
J
J
J
J
J
J
J
J
7334.59940
7361.79958
7364.82184
7599.00621
7602.02813
7679.08274
7682.10451
7941.97621
8088.53133
8306.10046
7362.72643
7587.95740
7600.06699
8079.58054
7981.98734
7986.03382
8054.87480
8008.03130
8313.02465
7966.01321
8012.04130
8048.86235
7998.88349
8097.72511
8283.05268
0.00062
0.00041
0.00076
0.00052
0.00036
0.00019
0.00022
0.00038
0.00056
0.00025
0.00034
0.00050
0.00047
0.00045
0.00085
0.00050
0.00048
0.00062
0.00035
0.00042
0.00081
0.00026
0.00056
0.00031
0.00093
0.696
0.710
0.717
0.707
0.698
0.733
0.703
0.648
0.691
0.725
0.653
0.625
0.612
0.606
0.499
0.316
0.408
0.522
0.555
0.722
0.572
0.807
0.746
0.774
0.771
0.074
0.063
0.065
0.063
0.046
0.037
0.034
0.049
0.074
0.042
0.037
0.056
0.070
0.106
0.058
0.039
0.044
36.56
37.38
35.68
36.34
37.00
36.87
34.46
35.49
35.67
36.02
41.69
41.94
42.20
42.28
48.99
47.88
50.18
0.052
0.034
56.03
56.80
0.046
0.053
0.038
0.048
0.037
0.111
61.76
62.08
62.04
69.96
70.57
69.14
1.38
1.11
1.84
1.20
1.00
0.61
0.72
0.43
1.28
0.38
1.00
0.98
1.24
1.16
0.42
1.33
1.28
0.42
1.04
1.29
0.42
0.39
1.51
1.10
0.43
0.17
0.18
0.26
0.17
0.15
0.15
0.28
0.32
0.23
0.20
0.23
0.19
0.17
0.18
0.22
0.23
0.15
0.20
0.18
0.41
0.31
0.35
0.32
0.29
0.37
0.12
0.11
0.16
0.12
0.11
0.10
0.12
0.13
0.14
0.12
0.13
0.12
0.11
0.11
0.12
0.13
0.10
0.12
0.11
0.10
0.12
0.11
0.13
0.12
0.12
3.4 Global analysis
For each planet, we ran a global MCMC analysis of all the
transits to check the consistency of the transits depths from
the individual MCMC analysis. We used only one light curve
from the set of all light curves selected for each transit obser-
vation. Said light curve's median value of the posterior PDF
for the transit depth was the closest to the median value ob-
tained from the combination of all the PDFs deduced from
the set of best light curves for a particular transit. These
light curves are presented in Appendix A1 along with cor-
responding RMS versus bin size plots made with the MC3
software (Cubillos et al. 2017) to assess the amount of corre-
lated (red) noise in our time-series. The global analysis was
performed for each planet in each filter -- TRAPPIST-1 b-g
in J band and TRAPPIST-1 b in NB2090 band. We note
that there is only one transit of planet c in NB2090 band.
Jump parameters were the same as for the individual
analyses (see Subsection 3.3), but with an addition of a Tran-
sit Timing Variation (TTV) for each transit. Period P and
initial transit epoch T0 were fixed to the values from Delrez
et al. (2018). Thus there were seven common parameters for
all transits (stellar parameters including LD coefficients +
transit impact parameter b) and two individual parameters
for each transit (dF and TTV). We used the same photo-
metric baselines that were applied in the individual analy-
ses. First, we ran MCMC with 10 000 steps to estimate the
MNRAS 000, 1 -- 21 (2019)
correction factors, CF, and then two chains of 100 000 steps
with 20% burn-in phase were executed to derive the transit
parameters. Their convergence was checked as well with the
test of Gelman & Rubin. Deduced transit depths are pre-
sented in Table 4 and detrended phase-folded light curves
for each planet and bandpass are presented in Fig. 3.
4 RESULTS AND DISCUSSION
4.1 Transit depths variability
To assess how transit depths change from epoch to epoch, for
each planet we compared the depths inferred from the indi-
vidual analyses with the one measured from the global anal-
ysis of all transits (see Table 3 and Table 4, respectively). We
conclude that the individual transit depths of TRAPPIST-
1 b, c, e, and g are consistent with the values deduced from
the global analysis at better than 1-σ, and that the maxi-
mum standard deviation of the measured individual transit
depth does not exceed 260 ppm (the case of TRAPPIST-1 b,
where the depth inferred from light curve #8 could deviate
the most because of a data gap during the egress and a
short OOT baseline which prevents us from properly mod-
elling the systematic effects during the observing run). We
note that our initial standard deviation of the measured in-
dividual transit depth of planet b, which was obtained using
8
A. Burdanov et al.
Figure 2. Transit depths measured individually for each transit.
Values are displayed chronologically, but not linearly in time. The
horizontal black line shows the median value of the transit depth
posterior PDF inferred from the global analysis with 1-σ and
2-σ intervals shown in shades of grey with numerical values on
the right. The dotted line represents the same transit depth from
the global Spitzer data analysis in the thermal-IR 4.5 µm range
(Delrez et al. 2018).
a classical approach to differential photometry by minimiza-
tion of the OOT RMS, was 1000 ppm. However, even with
a rigorous approach, transit depths of TRAPPIST-1 d and f
show temporal variations from epoch to epoch, with peak-
to-peak values of 1800 ppm, and 2400 ppm, respectively, and
standard deviations are larger than the mean errors.
We note that there are only three transits for planets d
and f, the two that appear to have noticeable temporal vari-
ations (Ducrot et al. (2018) also reports temporal variations
of planet d basing on 10 transits from K2 ). In the case of
planet d, all transit observations were done in photometric
conditions, but the first transit (light curve #15) has larger
error bars than the others, most likely because of a data gap
during an egress. Another transit of TRAPPIST-1 d suffered
from a short data gap in the bottom part (light curve #16).
In each case, after resuming the observations, the telescope
pointing held, holding all the stars within 1 pixel from their
initial positions, but gaps during the ingress/egress could
affect transit shapes, which could affect our deduced transit
depths. A similar situation can be seen during a transit of
planet b (above-mentioned light curve #8), where a data
gap is also present during an egress. However in the case
of a transit of planet g (light curve #25) data gap during
ingress and short OOT baseline are also present, but the
deduced transit depth is consistent with previous measure-
ments (light curves #23 and #24). In the case of planet f,
the first transit was observed in thin cirrus conditions (light
curve #20), the second transit shows a possible spot-crossing
event (light curve #21) and the third transit suffered from
thin cirrus before the ingress (light curve #22). All tran-
sits of TRAPPIST-1 d and f were observed with UKIRT in
J band with the same instrumental setup. As precipitable
water vapour (PWV) has drastic influence on the opacity of
Earth's atmosphere in the near-IR and thus impacts near-IR
photometry, we checked each observing night for the amount
of PWV6 and we could not find any correlations between
deduced transits depths and the amount of PWV. Besides
possible effect of thin cirrus clouds, we could not attribute
transit variations in depth to other external factors, such as
the position of the star on the chip, mean FWHM, time, air-
mass or sky background values or any combination of these.
Thus we conclude that for now our apparent high scatter
of transit depths for planets d and f likely originate from
the effects of the data gaps, thin cirrus or the spot-crossing
event.
Interestingly, all transits of TRAPPIST-1 d and f in
the near-IR were obtained in one observing season dur-
ing Summer and Autumn of 2017. Some of the transits
of TRAPPIST-1 b, c, e, and g were also obtained in the
same season, but their values are consistent with observa-
tions made in other seasons. In Fig. 4 we display the ratio
of an individual transit depth to the average transit depth
inferred from the global analysis for each planet as a func-
tion of Julian Date. If the scatter noticed for planets d and
f originated from a maximum of the star's magnetic cycle
in 2017, we would expect the transits of the other planets
to have been affected too. The fact that it was not the case
argues against this hypothesis.
All our observations were centered only on transit win-
dows with the longest observing window lasting 5 h which
makes it hard to properly sample the rotational period of the
star, reported to be ∼3.3 d basing on the K2 data (Luger
et al. 2017). However, this period could be a characteris-
tic timescale of active regions (Morris et al. 2018a). In any
case, more transit observations of TRAPPIST-1 planets, es-
pecially of planets d and f, and frequent photometric mon-
itoring of the host star are needed to confirm astrophysical
origin and understand the real cause of transit depth varia-
tions of TRAPPIST-1 d and f.
6 http://www.eao.hawaii.edu/weather/watervapor/mk/
archive/
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
9
Table 4. Transit depths from the global analyses: median values
of the posterior PDFs and their respective 1-σ limits.
Planet
TRAPPIST-1 b
TRAPPIST-1 c
TRAPPIST-1 d
TRAPPIST-1 e
TRAPPIST-1 f
TRAPPIST-1 g
dFJ
(per cent)
0.700 ± 0.023
0.641 ± 0.034
0.387 ± 0.029
0.550 ± 0.029
0.759 ± 0.031
0.758 ± 0.030
dFNB2090
(per cent)
0.688 ± 0.05
0.618 ± 0.04
-
-
-
-
Table 5. Standard deviations and mean errors of the transit
depths from the individual analyses.
Planet
Filter
σ
(per cent)
Mean error
(per cent)
J
TRAPPIST-1 b
TRAPPIST-1 b NB2090
TRAPPIST-1 c
TRAPPIST-1 d
TRAPPIST-1 e
TRAPPIST-1 f
TRAPPIST-1 g
J
J
J
J
J
0.026
0.004
0.021
0.091
0.023
0.119
0.015
0.050
0.074
0.054
0.047
0.043
0.046
0.066
Figure 3. Period-folded transits of TRAPPIST-1 b-g planets
multiplied by the baseline polynomials, and corrected for TTVs.
Individual measurements are presented in colored circles and
white circles are 7 min binned values. The solid black line repre-
sents the best-fit model.
4.2 Updated transmission spectra and stellar
contamination
We updated the transmission spectra of TRAPPIST-1 b-g
planets presented in Ducrot et al. (2018) by adding transit
depths deduced from our global analysis in J and NB2090
bands. Our data set only includes two NB2090 band tran-
sits of TRAPPIST-1 b and one of TRAPPIST-1 c which is a
cleaner, narrower bandpass (see Subsection 2.2). Additional
transit depth measurements are especially needed with this
MNRAS 000, 1 -- 21 (2019)
Figure 4. Ratio of the individual transit depth to the average
depth from the global MCMC analysis as a function of time (JD).
0.040.030.020.010.000.010.020.030.04Time since mid-transit (d)0.770.790.810.830.850.870.890.910.930.950.970.991.01Relative fluxTRAPPIST-1b, J bandTRAPPIST-1b, NB2090 bandTRAPPIST-1c, J bandTRAPPIST-1c, NB2090 bandTRAPPIST-1d, J bandTRAPPIST-1e, J bandTRAPPIST-1f, J bandTRAPPIST-1g, J band10
A. Burdanov et al.
bandpass to improve the precision. The updated transmis-
sion spectra is presented in Fig. 5.
We note that transit depths of planet b in the near-IR
are consistent with previously published non-HST transit
depth measurements. They lie in most cases within ∼ 1-σ
of the HST results, but are consistently shallower in depth.
HST measurements in their turn could be affected by the
telescope orbit-dependent systematic effects, which could re-
sult in much deeper transit depths (see de Wit et al. (2016),
Subsection 3.2 in Ducrot et al. (2018) and Zhang et al. (2018)
for details). For TRAPPIST-1 c our deduced transit depths
also disagree somewhat with HST measurements and show,
like TRAPPIST-1 b, shallower transits comparing to Spitzer
4.5 µm depths, but lie within the 3-σ interval. Transit depths
of TRAPPIST-1 d are in agreement with HST data, and
show excellent agreement for g. For planets e and f our de-
duced depths in 1.2 µm are deeper than all published depths
in other spectral ranges. But these findings are based on just
two transits of TRAPPIST-1 e, and transits of TRAPPIST-
1 f show considerable temporal variations.
Though there is a certain difficulty in obtaining ab-
solute near-IR transit depths, it may be mitigated by us-
ing different instruments with partially overlapping wave-
length bands. This mitigation procedure requires spec-
tra taken with different instruments requires simultaneous
multi-instruments observations in order to avoid compen-
sating for time- and wavelength-dependent effects that may
contaminate a planetary transmission spectrum, such as the
epoch-dependent brightness distribution of the host star.
We also added our depth measurements for planets b
and c to the stellar contamination spectra proposed by Mor-
ris et al. (2018a), which originate from the possible existence
of bright spots of temperature 4500 K. We find that the
prediction implies a flat contamination spectrum for wave-
lengths redder than 0.7 µm and that this model disagrees
with what we see in near-IR (see Fig. 6). Since the real distri-
bution of the spots and their temperatures are unknown, the
TRAPPIST-1 system would benefit from additional moni-
toring in the near-IR to put better constraints on contami-
nation models. In addition, current contamination models
only account for spots or faculaes or their combinations,
but the atmosphere of TRAPPIST-1 could contain dust.
Miles-P´aez et al. (2019) using linear polarization photom-
etry in the near-IR J band finds hints that the atmosphere
of TRAPPIST-1 is quite dusty. Likely, this should also be
included for proper accounting of stellar contamination.
5 CONCLUSIONS
We presented here an extensive photometric data set of
25 transits of TRAPPIST-1 observed in the near-IR with
UKIRT and AAT in J band (1.2 µm) and with the VLT in
the NB2090 band (2.1 µm) from 2015 to 2018. We deduced
individual transit depths for each transit taking into account
inherent to ground-based near-IR observations correlations
of transit parameters with the selected sets of comparison
stars and photometric aperture sizes to obtain results as ro-
bust and uniform as possible.
We reach a photometric precision of ∼ 0.003 (RMS of
the residuals), and we detect no significant temporal varia-
tions of transit depths of TRAPPIST-1 b, c, e, and g, while
Figure 5. Transmission spectra of TRAPPIST-1 b-h planets. The
solid black line is the weighted mean of all measurements exclud-
ing HST presented in Ducrot et al. (2018) with corresponding 1-σ
confidence intervals in shades of gray and numerical values on the
right. HST measurements are presented as grey points. Measure-
ments in the near-IR (1.2 µm and 2.1 µm) from this study are
presented as colored circles with white center. Each point is a
median value of the posterior PDF and its respective 1-σ from
the global analysis at the effective wavelength of corresponding
instrument.
transit depths of planets d and f show hints of variability
with peak-to-peak values of 1800 and 2400 ppm, respec-
tively. Besides thin cirrus observing conditions, we could
not link these depth variations to any external parameters,
including the amount of precipitable water vapour, the po-
sition of the star on the detector, the mean FWHM, etc.
We deduce that they likely originate from a few abnormal
transits that were affected by thin cirrus, data gaps or, for
MNRAS 000, 1 -- 21 (2019)
T1b 0.719 0.714 0.725T1c 0.692 0.687 0.698T1d 0.360 0.354 0.366T1e 0.481 0.473 0.489T1f 0.639 0.629 0.648T1g 0.760 0.750 0.770T1h 0.332 0.320 0.3441.02.5.Wavelength [µm]Transit Depth [%] . . . .Ground-based follow-up of TRAPPIST-1 in the near-IR
11
gique (F.R.S.-FNRS) under Grant No. 2.5020.11 and by the
Walloon Region.
PyRAF is a product of the Space Telescope Science In-
stitute, which is operated by AURA for NASA.
Authors would like to thank Mike Road and Mike Irwin
from CASU for their help with access to the UKIRT data at
WSA; Staff Scientists Drs. Tom Kerr and Watson Varricatt,
Telescope Operators Sam Benigni, Tim Carroll, Eric Moore,
and Michael Pohlen for scheduling and performing observa-
tions with the UKIRT ; NASA, University of Arizona, Uni-
versity of Hawai'i, Lockheed Martin, and the US Naval Ob-
servatory for support and granting time for the above noted
observing programs; Duncan Wright and Tayyaba Zafar for
performing observations with the AAT ; Evgeny Ivanov for
his advises on (C´ECI) computations.
The research leading to these results have received fund-
ing from the ARC grant for Concerted Research Actions,
financed by the Wallonia-Brussels Federation, and by the
NASA K2 Cycle 4 program.
V.V.G. is a F.R.S.-FNRS Research Associate, M.G. and
E.J. are F.R.S.-FNRS Senior Research Associates.
Authors would like to thank the anonymous reviewer
for the constructive comments, which helped us to improve
the quality of the paper.
REFERENCES
Apai D., et al., 2018, preprint, (arXiv:1803.08708)
Barstow J. K., Irwin P. G. J., 2016, Monthly Notices of the Royal
Astronomical Society: Letters, 461, L92
Casali M., et al., 2007, A&A, 467, 777
Claret A., Hauschildt P. H., Witte S., 2012, A&A, 546, A14
Clark B. J. M., Anderson D. R., Madhusudhan N., Hellier C.,
Smith A. M. S., Collier Cameron A., 2018, A&A, 615, A86
Croll B., et al., 2015, ApJ, 802, 28
Cubillos P., Harrington J., Loredo T. J., Lust N. B., Blecic J.,
Stemm M., 2017, AJ, 153, 3
Delrez L., et al., 2018, MNRAS, 475, 3577
Ducrot E., et al., 2018, AJ, 156, 218
Gelman A., Rubin D. B., 1992, Statistical Science, 7, 457
Gillon M., et al., 2012, A&A, 542, A4
Gillon M., et al., 2014, A&A, 563, A21
Gillon M., et al., 2016, Nature, 533, 221
Gillon M., et al., 2017, Nature, 542, 456
Gizis J. E., Monet D. G., Reid I. N., Kirkpatrick J. D., Liebert
J., Williams R. J., 2000, AJ, 120, 1085
Grimm S. L., et al., 2018, A&A, 613, A68
Howell S. B., 2006, Handbook of CCD Astronomy
Lendl M., Gillon M., Queloz D., Alonso R., Fumel A., Jehin E.,
Naef D., 2013, A&A, 552, A2
Lincowski A. P., Meadows V. S., Crisp D., Robinson T. D., Luger
R., Lustig-Yaeger J., Arney G. N., 2018, ApJ, 867, 76
Luger R., et al., 2017, Nature Astronomy, 1, 0129
Mandel K., Agol E., 2002, ApJ, 580, L171
Miles-P´aez P. A., Zapatero Osorio M. R., Pall´e E., Metchev S. A.,
2019, MNRAS, 484, L38
Moran S. E., Horst S. M., Batalha N. E., Lewis N. K., Wakeford
H. R., 2018, AJ, 156, 252
Morley C. V., Kreidberg L., Rustamkulov Z., Robinson T., Fort-
ney J. J., 2017, ApJ, 850, 121
Morris B. M., Agol E., Davenport J. R. A., Hawley S. L., 2018a,
ApJ, 857, 39
Morris B. M., et al., 2018b, ApJ, 863, L32
Rackham B. V., Apai D., Giampapa M. S., 2018, ApJ, 853, 122
Figure 6. Stellar contamination spectra proposed by Morris et al.
(2018a) originating from possible existence of bright spots of tem-
perature 4500 K (gray continuous line) with plotted observed
depth variations displayed as red points (depths from HST are
presented as grey points). The gray horizontal bars are the band-
integrated value for stellar contamination spectra where the inte-
grals are weighted uniformly in wavelength.
one, by a possible spot-crossing event. Considering the small
number of observed transits for TRAPPIST-1 d and f, the
system would benefit from more transit observations and
from frequent monitoring in the near-IR to probe photomet-
ric activity of the host star, to better characterize its photo-
spheric homogeneity and to understand if hints of temporal
variability are of astrophysical origin.
We did not detect any flare nor transit that could be
attributed to other undetected planets. One possible clear
spot crossing event of TRAPPIST-1 f is presented in light
curve #21 with an amplitude of 200 ppm. We also computed
transit timings which will be helpful for further mass and
densities updates of the planets via TTV studies. Our depth
measurements for planets b and c disagree with the stellar
contamination spectra proposed by Morris et al. (2018a),
which originate from the possible existence of bright spots of
temperature 4500 K. Finally, updated transmission spectra
are presented for the six inner planets of the system. We
conclude that spectra of TRAPPIST-1 b and g are globally
flat, but some structures are seen for planets c, d, e and f
what could be epoch-dependent and coming from the stellar
origin.
ACKNOWLEDGEMENTS
TRAPPIST-1 observations in NB2090 band are based
the European Southern
on observations collected at
Observatory under ESO programmes
296.C-5010(A)
and 598.C-0347(A). Observations with UKIRT were
taken
projects U/15B/NA05A,
U/16A/NA04, U/16B/NA03, U/17A/NA01, U/17B/NA01,
U/18A/NAV05, and U/18B/NAV03. Observations with
the AAT are based on programmes A/2016B/109 and
A/2017B/104.
non-survey
through
Computational resources have been provided by the
Consortium des ´Equipements de Calcul Intensif (C´ECI),
funded by the Fonds de la Recherche Scientifique de Bel-
MNRAS 000, 1 -- 21 (2019)
1.02.0.55.0.00.51.01.51.52.01.21.31.41.51.6 . . . .Wavelength [µm]Transit Depth [%]TRAPPIST−1 b+c12
A. Burdanov et al.
Rodler F., L´opez-Morales M., 2014, ApJ, 781, 54
Schwarz G., 1978, Annals of Statistics, 6, 461
Siebenmorgen R., Carraro G., Valenti E., Petr-Gotzens M., Bram-
mer G., Garcia E., Casali M., 2011, The Messenger, 144, 9
Tinney C. G., et al., 2004, in Moorwood A. F. M., Iye M., eds,
Proc. SPIEVol. 5492, Ground-based Instrumentation for As-
tronomy. pp 998 -- 1009, doi:10.1117/12.550980
Tody D., 1986, in Crawford D. L., ed., Proceedings of the Meet-
ing, Tucson, AZ, March 4-8, 1986 Vol. 627, Instrumentation
in astronomy VI. Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference Series, Bellingham, WA, p. 733
Van Grootel V., et al., 2018, ApJ, 853, 30
Wakeford H. R., et al., 2019, AJ, 157, 11
Winn J. N., et al., 2008, ApJ, 683, 1076
Zhang Z., Zhou Y., Rackham B., Apai D., 2018, preprint,
(arXiv:1802.02086)
de Wit J., et al., 2016, Nature, 537, 69
de Wit J., et al., 2018, Nature Astronomy, 2, 214
APPENDIX A: LIGHT CURVES AND
RESPECTIVE BINNED RESIDUALS RMS
VERSUS BIN SIZE PLOTS
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
13
MNRAS 000, 1 -- 21 (2019)
Figure A1. light curves (LC) #1-6. See Tables 1,2, and 3 for the details.
14
A. Burdanov et al.
Figure A2. light curves (LC) #7-12. See Tables 1,2, and 3 for the details. MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
15
MNRAS 000, 1 -- 21 (2019)
Figure A3. light curves (LC) #13-18. See Tables 1,2, and 3 for the details.
16
A. Burdanov et al.
Figure A4. light curves (LC) #19-24. See Tables 1,2, and 3 for the details. MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
17
Figure A5. light curve (LC) #25. See Tables 1,2, and 3 for the details.
MNRAS 000, 1 -- 21 (2019)
18
A. Burdanov et al.
Figure A6. Binned residuals RMS versus bin size plots for light curves (LC) #1-12.
MNRAS 000, 1 -- 21 (2019)
Ground-based follow-up of TRAPPIST-1 in the near-IR
19
Figure A7. Binned residuals RMS versus bin size plots for light curves (LC) #13-24.
MNRAS 000, 1 -- 21 (2019)
20
A. Burdanov et al.
Figure A8. Binned residuals RMS versus bin size plots for light curves (LC) #25.
MNRAS 000, 1 -- 21 (2019)
This paper has been typeset from a TEX/LATEX file prepared by the author.
Ground-based follow-up of TRAPPIST-1 in the near-IR
21
MNRAS 000, 1 -- 21 (2019)
|
1706.06609 | 1 | 1706 | 2017-06-20T18:07:44 | Evaluating the Dynamical Stability of Outer Solar System Objects in the Presence of Planet Nine | [
"astro-ph.EP"
] | We evaluate the dynamical stability of a selection of outer solar system objects in the presence of the proposed new Solar System member Planet Nine. We use a Monte Carlo suite of numerical N-body integrations to construct a variety of orbital elements of the new planet and evaluate the dynamical stability of eight Trans-Neptunian objects (TNOs) in the presence of Planet Nine. These simulations show that some combinations of orbital elements ($a,e$) result in Planet Nine acting as a stabilizing influence on the TNOs, which can otherwise be destabilized by interactions with Neptune. These simulations also suggest that some TNOs transition between several different mean-motion resonances during their lifetimes while still retaining approximate apsidal anti-alignment with Planet Nine. This behavior suggests that remaining in one particular orbit is not a requirement for orbital stability. As one product of our simulations, we present an {\it a posteriori} probability distribution for the semi-major axis and eccentricity of the proposed Planet Nine based on TNO stability. This result thus provides additional evidence that supports the existence of this proposed planet. We also predict that TNOs can be grouped into multiple populations of objects that interact with Planet Nine in different ways: one population may contain objects like Sedna and 2012 VP$_{113}$, which do not migrate significantly in semi-major axis in the presence of Planet Nine and tend to stay in the same resonance; another population may contain objects like 2007 TG$_{422}$ and 2013 RF$_{98}$, which may both migrate and transition between different resonances. | astro-ph.EP | astro-ph | Draft version April 29, 2019
Preprint typeset using LATEX style emulateapj v. 12/16/11
7
1
0
2
n
u
J
0
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
0
6
6
0
.
6
0
7
1
:
v
i
X
r
a
EVALUATING THE DYNAMICAL STABILITY OF OUTER SOLAR SYSTEM OBJECTS
IN THE PRESENCE OF PLANET NINE
Juliette C. Becker1, Fred C. Adams1,2, Tali Khain2, Stephanie J. Hamilton2, David Gerdes1,2
Draft version April 29, 2019
ABSTRACT
We evaluate the dynamical stability of a selection of outer solar system objects in the presence
of the proposed new Solar System member Planet Nine. We use a Monte Carlo suite of numerical
N-body integrations to construct a variety of orbital elements of the new planet and evaluate the
dynamical stability of eight Trans-Neptunian objects (TNOs) in the presence of Planet Nine. These
simulations show that some combinations of orbital elements (a, e) result in Planet Nine acting as a
stabilizing influence on the TNOs, which can otherwise be destabilized by interactions with Neptune.
These simulations also suggest that some TNOs transition between several different mean-motion
resonances during their lifetimes while still retaining approximate apsidal anti-alignment with Planet
Nine. This behavior suggests that remaining in one particular orbit is not a requirement for orbital
stability. As one product of our simulations, we present an a posteriori probability distribution for
the semi-major axis and eccentricity of the proposed Planet Nine based on TNO stability. This result
thus provides additional evidence that supports the existence of this proposed planet. We also predict
that TNOs can be grouped into multiple populations of objects that interact with Planet Nine in
different ways: one population may contain objects like Sedna and 2012 VP113, which do not migrate
significantly in semi-major axis in the presence of Planet Nine and tend to stay in the same resonance;
another population may contain objects like 2007 TG422 and 2013 RF98, which may both migrate and
transition between different resonances.
1.
INTRODUCTION
In our solar system, a large population of small, rocky
objects resides beyond the orbit of Neptune, and the
collective structure of this population is anomalous, ex-
hibiting trends unexplained by random chance. Many
of these objects appear to occupy a region close to the
plane containing the eight known planets, leading to this
region being called the Kuiper Belt. The existence of
these objects has implications for the formation mecha-
nism of our solar system; however, we have yet discovered
only a small fraction of the objects orbiting beyond Nep-
tune. Since the turn of the century, many new objects
have been discovered in the Kuiper Belt. The subset
of objects orbiting outside of Neptune's orbit are called
Trans-Neptunian objects (TNOs), and they often have
dynamically interesting orbits.
In particular, some of
these objects have large semi-major axes and large per-
ihelion distances, including, for example, Sedna (Brown
et al. 2004), 2004 VN112 (Becker et al. 2008), 2010 GB174
(Chen et al. 2013), and 2012 VP113 (Trujillo & Sheppard
2014).
When Trujillo & Sheppard (2014) reported the discov-
ery of 2012 VP113, they also noted a curious clustering
in argument of perihelion for the population of TNOs
with high-a, high-q orbits. The authors proposed that
the high-q orbits could be generated in three ways: first,
by the ejection of a solar system body that left behind
the observed clustering as a signature of its ejection; sec-
ond, through a stellar fly-by encounter that perturbed
the orbits of some TNOs into their current configura-
[email protected]
1 Astronomy Department, University of Michigan, Ann Arbor,
MI 48109, USA
2 Physics Department, University of Michigan, Ann Arbor, MI
48109, USA
tions (Morbidelli & Levison 2004), where such interac-
tions are relatively common in the birth cluster (Li &
Adams 2015); third, through the presence of an addi-
tional planet in the solar system. Notably, this third
mechanism could also explain the clustering of this pop-
ulation of TNOs, as the proposed ninth planet's repeated
secular interactions with the shorter-period TNOs could
force the TNOs to keep their ω confined to be near either
0 degrees or 180 degrees (for example, this could occur
via the Kozai mechanism).
Batygin & Brown (2016a) also suggested the existence
of an additional planet (Planet Nine), which is differen-
tiated from the potential planet discussed in Trujillo &
Sheppard (2014) in the way it interacts with the TNOs.
The Batygin & Brown (2016a) version of Planet Nine
functions by explaining both the apsidal and ascending
node alignment of a selection of objects in the Kuiper
Belt. The objects under consideration in Batygin &
Brown (2016a) were two overlapping sets of objects: first,
those that have perihelion distance q > 30 AU and semi-
major axis a > 150 AU, while being dynamically stable
in the presence of Neptune; second, any objects which
have q > 30 AU and semi-major axis a > 250 AU, all
of which exhibit clustering in (when is defined as
the longitude of perihelion, = ω + Ω). The orbit for
Planet Nine presented in Batygin & Brown (2016a) was
a rough estimate, with semi-major axis a = 700 AU,
eccentricity e = 0.6, inclination i = 30 degrees, longi-
tude of ascending node Ω = 113 degrees, and argument
of perihelion ω = 150 degrees. The authors estimated
the planet to have a mass of roughly 10 Earth masses,
but all of these orbit predictions were noted as approxi-
mate. This mass estimate was supported in a follow-up
effort by the predicting authors (Brown & Batygin 2016),
which placed constraints on the orbital elements of the
2
Becker et al.
potential Planet Nine by using N-body simulations to de-
termine which Planet Nine realizations lead to clustered
TNOs in simulations. In this work, the authors deter-
mined that a 10 Earth mass Planet Nine was more likely
to recreate the observed clustering than a 20 Earth mass
Planet Nine.
The possibility of a new planet has led to a great deal
of recent work, e.g., to evaluate the likelihood of Planet
Nine's existence given the known properties of the Solar
System. For example, Fienga et al. (2016) and Holman
& Payne (2016) examined the measured Earth-Saturn
distance, and determined that Planet Nine is likely to
have a true anomaly near 117 degrees (based on the
upper limit to the amplitude of the residuals for that
distance measurement).
In complementary work, Mal-
hotra et al. (2016) evaluated the potential resonant be-
havior that Planet Nine could invoke in the population
of long-period TNOs, and predicted that Planet Nine
should have a semi-major axis of a = 665 AU in order
to support stability-boosting resonances. Other authors
(Lawler et al. 2016; de la Fuente Marcos et al. 2016) have
examined the dynamical effects that Planet Nine would
have on the populations of objects that exist in the So-
lar System. Finally, additional work was carried out to
explain how Planet Nine would fit into our existing pic-
ture of the Solar System. For example, multiple groups
found that the existence of such a ninth planet can be in-
voked to explain the six-degree obliquity of the sun (Bai-
ley et al. 2016; Lai 2016; Gomes et al. 2017). Several pre-
vious works have also suggested that the mechanism by
which Planet Nine shapes the orbits of the TNOs may be
orbital resonances (Batygin & Brown 2016a; Beust 2016;
Malhotra et al. 2016; Millholland & Laughlin 2017).
The conclusion in Batygin & Brown (2016a) was based
only on six of the (at the time) thirteen discovered ex-
treme TNOs. These six TNOs (2004 VN112, 2007 TG422,
2010 GB174, 2012 VP113, 2013 RF98, and Sedna) were
chosen because they exhibit clustering in and have
a > 250 AU, meaning they are expected to be influ-
enced by Planet Nine. Sheppard & Trujillo (2016) also
announced the discovery of two new objects (2014 SR349
and 2013 FT28) that also fit into the class of objects used
by Batygin & Brown (2016a) to predict the existence of
Planet Nine. In order for Planet Nine to be capable of
forcing the orbits of the TNOs into aligned configura-
tions, the TNOs must remain dynamically stable on sec-
ular timescales, to allow apsidal alignment to occur. Ob-
jects whose orbital elements are affected by short-period
scattering events can be used to understand the scatter-
ing event itself, but not the long-period, secular effects in
the system. Numerical simulations in Batygin & Brown
(2016a) and Sheppard & Trujillo (2016) determined that
some of the high semi-major axis (a > 250) objects' or-
bits change significantly over 1 Gyr timescales. They
used numerical simulations to evaluate the stability of
these bodies in the presence of the four giant planets, and
found that some objects were not dynamically stable. If
two of the six objects used to infer the existence of Planet
Nine are dynamically unstable (for example, being sus-
ceptible to scattering events) in the presence of Neptune,
it is less likely that those objects could reside in their cur-
rent orbits long enough to be influenced by Planet Nine
and become apsidally aligned via that mechanism. de
la Fuente Marcos et al. (2016) also found that the six
aforementioned extreme TNOs can become dynamically
unstable on relatively short timescales in the presence of
both Neptune and the nominal Planet Nine, which could
potentially prevent the observed apsidal alignment from
occurring on secular timescales.
Of course, dynamical
stability is a function of
timescale. The objects with perihelion distances in the
range 30-40 AU can be termed a part of the scattered disk
(Lykawka & Mukai 2007). The objects in this population
are characterized by repeated (potentially scattering) in-
teractions with Neptune (Nesvorn´y & Roig 2001). An in-
tegration of the solar system run for an indefinite amount
of time (without the presence of Planet Nine) will even-
tually lead to all objects in the scattered disk leaving the
solar system. However, our solar system is only 4.5 Gyr
old, and it does not take this long for Planet Nine to
align the TNOs into the pattern reported in Batygin &
Brown (2016a). For this reason, 'dynamical stability' in
this work refers to objects which remain in orbits com-
parable to their current orbits for 4.5 Gyr.
Sheppard & Trujillo (2016) reported two new TNOs in
the regime of interest for consideration of Planet Nine,
and the population of discovered TNOs and Kuiper Belt
Objects is rapidly growing. The Canada-France Ecliptic
Plane Survey (Petit et al. 2011), which ran from 2003 --
2007, detected 169 TNOs with a preference for larger
TNOs (size R > 100 km). A 32-square-degree survey
running from 2011 -- 2012 by Alexandersen et al. (2014)
detected 77 TNOs. The Outer Solar System Origins Sur-
vey (Bannister et al. 2016), which is currently in progress,
has thus far detected 85 TNOs. The Panoramic Survey
Telescope and Rapid Response System (Pan-STARRS) is
intended to discover comets and asteroids (particularly
near-Earth objects), but has also found Kuiper Belt Ob-
jects (Chen et al. 2016) and giant planet Trojans (Horner
et al. 2012; Guan et al. 2012; Lin et al. 2016). In addi-
tion to these dedicated Solar System searches, cosmolog-
ical searches also allow for the serendipitous discovery of
foreground TNOs. For example, the Dark Energy Survey
has experienced great success in discovering TNOs and
other solar system objects (Gerdes et al. 2016; Dark En-
ergy Survey Collaboration et al. 2016), including a new
dwarf planet (Gerdes et al. 2017).
This ever-growing population of TNOs will allow for
increasingly stronger constraints on the possible orbital
elements of Planet Nine and its current location in the
Solar System. These objects will also provide a clearer
picture of the dynamical regimes of bodies in the outer
Solar System, where these orbits are sculpted by Neptune
on the inside and could be influenced by Planet Nine from
the outside.
In this paper, we use as our sample the TNOs with
a > 250 AU from Batygin & Brown 2016a and two newly
discovered objects (announced in Sheppard & Trujillo
2016) expected to be dynamically stable in the presence
of the solar system objects. We use N-body techniques
to simulate the behavior of these objects in the presence
of Planet Nine to place limits on the possible orbital ele-
ments (a, e) of Planet Nine. In Section 2, we describe the
sets of simulations carried out in this work and present
some results. Our treatment uses a Monte Carlo sample
of 1500 Planet Nine realizations, and thus extends most
previous work, which generally considered only a single
nominal orbit or small number (generally N ≤ 3) of po-
Planet Nine Stability Survey
3
tential orbits. The simulations enable us to estimate the
mean lifetime of each TNO in the presence of each of the
Planet Nine realizations under consideration. In Section
3, we develop and provide a posterior probability dis-
tribution for the most likely values (a, e) for the Planet
Nine orbit, given that we observe the eight TNOs in their
current orbits. In Section 4, we discuss the different dy-
namical instability mechanisms contributing to the po-
tentially shortened lifetimes for the TNOs, and identify
some interesting differences between objects in our sam-
ple. In Section 5, we explore the relationship between the
period of our injected Planet Nine and the period of each
TNO, and show that TNOs living in dynamically stable
configurations tend to attain integer period ratios with
Planet Nine. Remarkably, some TNOs do not remain
in the same commensurability for their entire lifetimes;
instead, they transition between multiple near-resonant
locations. The paper concludes, in Section 6, with a sum-
mary of our main results and a discussion of avenues for
future study.
2. NUMERICAL SIMULATIONS OF TNO ORBITAL
EVOLUTION IN THE PRESENCE OF PLANET NINE
Brown & Batygin (2016) provided a relative poste-
rior probability distribution for the (a, e) of Planet Nine.
This posterior was constructed using clustering argu-
ments of the same type that were used in Batygin &
Brown (2016a) to predict the existence of Planet Nine.
The logic used in both these works can be summarized as
follows: the observed TNOs' orbits are aligned in physi-
cal space (the longitudes of perihelion and longitudes
of ascending node Ω are confined to a narrow range in
angles instead of uniformly populating all allowed an-
gles), and the probability of this occurring by chance is
low, even accounting for potential bias in the observa-
tions. Numerical N-body simulations of a selection of
test-particles at varying orbital radii do not recreate the
observed clustering unless Planet Nine is included in the
simulations; even then, some orbits of Planet Nine are
more likely to recreate the observed clustering than oth-
ers. From a suite of numerical integrations, Brown &
Batygin (2016) predicted the combinations of (a, e) that
were most likely to allow a population of test particles
that exhibited the observed clustering in longitude of per-
ihelion.
The physical orientation of orbits is not the only ob-
served physical property that can be measured from the
known TNOs. The most basic property that the ob-
served TNOs share is orbital stability: although merely
seeing the TNO orbits today does not ensure that they
are dynamically stable in those current orbits, the ob-
served (Trujillo & Sheppard 2014; Batygin & Brown
2016a) physical clustering of the orbit directions suggests
that the orbits have been dynamically stable for a long
enough time for this alignment to develop. Whether or
not the aligning agent is the theorized Planet Nine, this
alignment would have to have taken place over secular
timescales, suggesting that these TNOs must have been
dynamically stable over such timescales.
The dynamical stability of these objects is thus a fun-
damental and necessary property. As such, any allowable
Planet Nine would have to allow the continued dynami-
cal stability of the TNOs over secular and solar system
lifetimes. To evaluate the likelihood of any particular re-
alization of Planet Nine, we must evaluate the lifetimes
of the TNOs in the presence of said Planet Nine.
2.1. Numerical Methods
To determine how the lifetimes of the TNOs vary in
the presence of different combinations of the semi-major
axis and eccentricity of Planet Nine, we run a large num-
ber of numerical N-body integrations, each including one
potential realization of Planet Nine and the population
of TNOs we are testing. This population includes the
six TNOs considered in Batygin & Brown (2016a) (2004
VN112, 2007 TG422, 2010 GB174, 2012 VP113, 2013 RF98,
and Sedna) and two additional objects discovered by
Sheppard & Trujillo (2016) (2014 SR349 and 2013 FT28)
that appear to fit in the same dynamical class as the pre-
vious six. The orbital elements of all eight objects in our
sample are reported in Table 1. The objects we consider
in this work are those with semi-major axis a > 250 AU
and perihelion distance q > 30 AU. This is a different
sample, with a stricter semi-major axis cut, than was
originally used in Trujillo & Sheppard (2014) to iden-
tify the clustering effect (Trujillo & Sheppard 2014, used
a > 150 AU). We limited the objects considered in this
paper to those with a > 250 AU for three reasons: first,
these are the objects which are found in Batygin & Brown
(2016a) to exhibit both confinement in the longitude of
perihelion and in the longitude of ascending node; sec-
ond, these large-a objects exhibit a variability in semi-
major axis (Batygin & Brown 2016a) that can poten-
tially lead to dynamical instability; third, we expect the
dynamics of these long period orbits to be dominated by
Planet Nine rather than by Neptune (Sheppard & Tru-
jillo 2016).
We choose to explore the effect of Planet Nine's a and
e because the orbital angles of Planet Nine are directly
oppositional to those of (all but one of) the discovered
ETNOs, and the current location of Planet Nine can
be estimated using other methods (for example, Fienga
et al. 2016; Holman & Payne 2016). For this work, the
inclination and orbital angles of Planet Nine were taken
to be the nominal values presented in Batygin & Brown
(2016a). The robust consideration of a wider range of
angles is beyond the scope of this work, but we report a
basic reproduction of the experiments of this work while
allowing the orbital angles to vary in Appendix B. We
choose to fix these angles for the remainder of the main
manuscript for two reasons: first, to examine the effect
of (a, e), we take a cross section in the other angles in
order to remove the degeneracies that would be imposed
by allowing these angles to vary; second, considering all
orbital angles would be too computationally intensive, so
we choose the best values we can and proceed under the
assumption that altering these angles will not change the
broad trends of the results.
To evaluate the dynamical stability (and thus, ex-
pected mean orbital lifetimes) of these eight TNOs in the
presence of Planet Nine, we ran 1500 numerical N-body
integrations using the hybrid symplectic and Bulirsch-
Stoer (B-S) integrator built into Mercury6 (Chambers
1999), and conserved energy to 1 part in 1010. We re-
placed the three inner giant planets (Jupiter, Saturn and
Uranus) with a solar J2 (as done in Batygin & Brown
2016a). We included Neptune as an active particle be-
cause the orbital motion of Neptune may induce scatter-
4
Becker et al.
ing or resonant effects on the TNOs, potentially leading
to dynamical instabilities (rapid, drastic orbital evolu-
tion for the TNOs). For each realization, we included
each TNO with orbital elements drawn from observa-
tional constraints, sampling each orbital element from
reported 1σ errors. Then, we integrated each of the 1500
realizations forward for 4.5 Gyr using computational re-
sources provided by Towns et al. (2014). For data man-
agement, we used the pandas python package (McKinney
2010).
Some of the TNOs (2013 RF98, for example) have very
rough observational priors. Since we are sampling from
observational priors in the orbital elements we assign to
these small bodies, the lifetimes of these objects (com-
pared to Sedna, whose orbital properties are much better
constrained) include the degeneracy of both Planet Nine
and also of the TNO itself. This leads to smoothed dis-
tributions, which are effectively convolutions of the true
lifetimes with an uncertainty kernel including the errors
of the measured orbital properties of the object.
The final result of these simulations is 1500 measures
of dynamical lifetime for each object, which can be plot-
ted on a (a, e) grid as shown in the top panel of Figure
1 (which shows the points for 2012 VP113). We do not
expect the lifetimes to form a smooth function for several
reasons: first, for each object, we draw orbital parame-
ters from observational priors, resulting in some expected
scatter in results for even a single Planet Nine realization;
second, chaotic effects will cause scatter in outcomes be-
tween realizations. For that reason, we must run enough
simulations that we can treat the averaged value from all
results near a Planet Nine (a, e) point as a good average
of the behavior that a given Planet Nine would engender.
We construct the contour plots in Figure 2 and Fig-
ure 4 by using a polyharmonic spline (from Jones et al.
2001) to smooth and interpolate between the points gen-
erated by our Monte Carlo simulations. We investigated
the effects of different interpolation methods, and found
that for our sample size (N ∼ 1500 realizations), the in-
terpolation scheme does not drastically affect the final
result. However, it is important to note that when we
attempted to use N ∼ 100 realizations, the results var-
ied drastically between different interpolation schemes.
Since our sample size does not show such variations be-
tween methods, we consider the (a, e) parameter space
well-sampled. Figure 1 demonstrates the smoothed two-
dimensional lifetime function and provides a comparison
to the raw points for one object (2012 VP113) in our sam-
ple. Analogous plots (not presented) can be constructed
for each of the eight objects in our sample.
2.2. Numerical Results
The results of the simulations are plotted in Figure
2, which presents contour plots for the expected orbital
lifetimes of the eight TNOs in the presence of various
realizations of Planet Nine. The main results that we get
from this experiment fit into three categories: we affirm
the potential stability of all eight TNOs in the presence
of Neptune and Planet Nine, we find that different TNOs
prefer different parameter spaces of Planet Nine, and we
examine the fate of the TNOs that do go dynamically
unstable.
Stability of TNOs. For all objects in our sample,
there are realizations of Planet Nine that allow them to
Fig. 1. -- (Top panel) The points for one object in our sample,
2012 VP113, which are color coded by the amount of time the
system remained dynamically stable. (Bottom panel) The same
points, turned into a contour plot by use of a polyharmonic spline.
The contour plot allows for easier comparison of the posterior for
each object, and visualizes the interpolation up for the Monte Carlo
sampling.
remain dynamically stable for the lifetime of the solar
system. Here, dynamical stability requires that a TNO
remain within 100 AU of its starting orbit, and not ex-
perience collisions with the planets or central body. This
value of 100 AU was chosen from examination of the
time-evolution of the bulk set of TNOs. We find that
when orbits vary by less than 100 AU, the objects are
generally confined to a network of mean motion reso-
nances (MMRs), an outcome that we describe in Section
5.
When orbits change by more than 100 AU, they po-
tentially change dynamical class. The criteria in Batygin
& Brown (2016a) included the cut that aT N O > 250 AU,
polyharmonic spline Planet Nine Stability Survey
5
Fig. 2. -- Individual contour plots of the lifetimes of each of the eight TNOs considered. The lifetime plots do not exhibit the same general
trends, indicating that the different objects may be members of different dynamical classes of objects. Table 1 presents the complete list
of their orbital properties; the title of each subplot here is the short name of the object it portrays.
and 100 AU of migration inwards would change this dy-
namical classification of 5 of the 8 TNOs considered in
this work. Since this choice of cutoff is somewhat arbi-
trary, we present in Table 2 the frequency of final dynam-
ical outcome for each TNO in our sample. Migration is
in outcome in between 7%-30%, depending on the TNO
under consideration, so future work that uses a different
criterion for migratory instability can expect variation
from our results of this magnitude.
Table 1 provides a 'Percent Stable' column, which gives
for each object the percentage of our 1500 realizations
(which all have different a, e for Planet Nine) which al-
lowed that object to remain dynamically stable for the
entire 4.5 Gyr simulation. Since our simulations also
include Neptune as an active particle, all of these ob-
jects can be dynamically stable in the presence of both
Neptune and Planet Nine. Notably, 18 of our 1500 real-
izations of Planet Nine allowed all eight TNOs to remain
dynamically stable for 4.5 Gyr. The 18 trials that allowed
all tested TNOs to survive had semi-major axes varying
between 600 AU and 800 AU and eccentricities between
0.35 and 0.55. The small sample (N ∼ 18) that allowed
all TNOs to survive limits any further conclusions based
on their orbital parameters. Sheppard & Trujillo (2016)
found that 2007 TG422 and 2013 RF98 were both dynam-
ically unstable in the presence of Neptune, but they did
not include a potential Planet Nine in their integrations.
We find that 2007 TG422 and 2013 RF98 have stability
percentages of less than 20%, making them less stable on
average than, for example, 2012 VP113. However, with a
dynamically favorable realization of Planet Nine, both of
these objects can remain dynamically stable for a solar
system lifetime. This result, combined with the result
from Sheppard & Trujillo (2016) that these two objects
are not dynamically stable in the presence of Neptune
alone, suggests that Planet Nine can boost the orbital
stability of these objects.
Variations between behavior of different TNOs.
For each object in our sample, different realizations of
Planet Nine lead to differing object lifetimes. This is
shown in Figure 2, which plots the lifetime of each TNO
as a function of the semi-major axis and eccentricity of
Planet Nine. These stability maps look very different
for different objects. For example, large-a, low-e orbits
lead to a longer lifetime for 2012 VP113. For this object,
shorter object lifetimes occur if Planet Nine has a shorter
perihelion distance (lower a or higher e). This is intu-
itive. In contrast, 2007 TG422 is dynamically unstable in
the presence of these same long perihelion-distance ob-
jects that 2012 VP113 preferred. Since different objects
prefer different regions of Planet Nine's possible parame-
ter space, a better understanding of the constraints they
provide can be obtained by considering all objects simul-
taneously. We will discuss how this can be done, as well
as find constraints using our results thus far, in Section
3.
Fate of the TNOs. To construct the lifetime con-
tours in Figure 2, we defined the lifetime of an object
as the length of time it lived in our simulation without
significant alterations in its orbit. We defined significant
alterations in orbit to be any of the following: (1) mi-
gration in semi-major axis by more than 100 AU from
the starting orbit of the object, (2) collisions or a close
encounter (defined as passing within 3 Hill radii of the
larger body) with Planet Nine or Neptune, (3) collision
with the central body, and (4) ejection from the solar sys-
tem. It is clear that each of these criteria have different
thresholds for importance, as well as different outcomes
for the object. In particular, while ending methods (2-4)
result in a violent end for the TNO, method (1) does not
necessarily remove the TNO from the solar system: in-
stead, the TNO's orbit is significantly altered, but the
TNO may continue to be a part of the solar system,
merely in a different location. Numerical simulations al-
low these effects to be disentangled.
In particular, we
find that the two objects (2007 TG422 and 2013 RF98)
that do not prefer the high-a, low-e realizations of Planet
Nine tend to exhibit the migratory end in those realiza-
50070090011000.10.30.50.70.9Eccentricity of Planet NineGB17450070090011000.10.30.50.70.9RF9850070090011000.10.30.50.70.9SEDNA50070090011000.10.30.50.70.9TG4225007009001100Semi-major axis of Planet Nine (AU)0.10.30.50.70.9Eccentricity of Planet NineVN1125007009001100Semi-major axis of Planet Nine (AU)0.10.30.50.70.9VP1135007009001100Semi-major axis of Planet Nine (AU)0.10.30.50.70.9SR3495007009001100Semi-major axis of Planet Nine (AU)0.10.30.50.70.9FT280.00.51.01.52.02.53.03.54.04.5Lifetime (Gyr)6
Becker et al.
TABLE 1
Stability of ETNOs in presence of Neptune and Planet Nine
Object Name
a (AU)
e
i (degrees) ω (degrees) Ω (degrees) Percent Stable
2003 VB12 (Sedna)
2004 VN112
2007 TG422
2010 GB174
2012 VP113
2013 RF98
2013 FT28
2014 SR349
499
318
482
371
261
325
310
288
0.85
0.85
0.93
0.87
0.69
0.88
0.86
0.84
11.92
25.56
18.59
21.54
24.06
29.61
17.3
18.0
311.5
327.1
285.8
347.8
292.8
316.5
40.2
341.3
144.5
66.0
113
130.6
90.8
67.6
217.8
34.8
45.3%
28.4%
14.6%
17.9%
56.5%
18.7%
19.9%
18.1%
¯δt1
300 Myr
625 Myr
1.24 Gyr
1.03 Gyr
9 Myr
1.01 Gyr
709 Myr
917 Myr
Note. -- A list of the TNOs used for the dynamical survey. The dynamical stability of each object was
evaluated using a suite of numerical N-body simulations. The orbital elements of each object provided in this
table are the best fit observational values. In the simulations, the orbital elements were drawn from the 1σ
distribution for each realization of each object. Also provided in this table is the percentage of realizations of
each object that are dynamically stable. It is important to note that this percentage reported (which is the
percentage of realizations of each object that are dynamically stable) is marginalized over all realizations of
Planet Nine included in the simulations, and thus the exact percentage is not meaningful. What is meaningful
is that the percentages are all non-zero, indicating that for a selection of orbital parameters and Planet Nine
realizations, each object in our sample can be dynamically stable. 1: ¯δt is the difference in median lifetime (over
all Planet Nine realizations) between two cases: case one being when the definition of dynamical instability
does not include migration, and case two being when migration of more than 100 AU constitutes dynamical
instability. Larger values indicate that TNOs are susceptible to significant (δa > 100) migrations in semi-major
axis.
tions, rather than experiencing a violent instability. We
will further discuss this effect in Section 4.
3. DERIVING CONSTRAINTS ON THE ORBITAL
ELEMENTS OF PLANET NINE
Given the predicted lifetime distributions constructed
in the previous section, we can constrain the orbital prop-
erties of Planet Nine by determining which (a, e) combi-
nations allow the continued dynamical stability of the
observed clustered TNOs. We make an assumption here
that in order for the TNOs to attain their observed
clustering, they must have remained in their currently
observed orbits for at least secular timescales. This
would require those orbits being dynamically stable in
the presence of Planet Nine. The process of constructing
these constraints will be detailed in this section.
3.1. Bayesian inference towards a posterior probability
distribution for the orbital elements of Planet Nine
The ultimate goal of our dynamical stability analysis
is to determine the posterior probability distribution for
the orbital elements (a, e) of Planet Nine. This will serve
as a check and supplement to previous orbital posteriors
in the literature, which were estimated using different
techniques. We define A to be the orbital elements of
Planet Nine, Ti to be the expected dynamical lifetime of
the ith TNO, and I to be all prior information that we
can incorporate into our models (for example: the fact
that Planet Nine exists is a prior we impose, as are the
observational errors of the discovered TNOs, which lead
to uncertainties on their orbits). With these definitions,
the property we wish to measure is P (ATi, I), the prob-
ability of A (Planet Nine's orbital elements), given the
lifetimes of the observed TNOs and our other knowledge.
This posterior probability function can be represented
using Bayes' Theorem as follows:
However, upon inspection of this expression, there is one
clear problem: P (Ti, IA) requires knowledge of Ti, the
actual lifetimes of the observed TNOs in our solar sys-
tem. This is not a property that can be measured. As
an added difficulty, the prior information encapsulated
in I includes large uncertainties on the orbital elements
of the observed TNOs.
To overcome this difficulty, we define a new parame-
ter, Di. Di is defined to be the computed lifetimes of
the observed TNOs, where these lifetimes are marginal-
ized over the observational priors of the TNOs' orbital
elements and our other uncertainties that were before
folded into I. Although we cannot measure the true life-
times Ti, we have constructed TNO lifetime estimates Di
as they depend on our priors by using numerical N-body
simulations. These simulations are described in the pre-
vious section, and a visualization of the derived lifetimes
is presented in Figure 2.
Now, we can rewrite the Bayesian statement of our
posterior in terms of this new variable:
P (DiA)P (A)
P (ADi) =
P (Di)
(2)
when P (DiA) is the probability of getting the numer-
ically measured lifetimes conditional on the orbital ele-
ments of Planet Nine, P (A) is our priors on Planet Nine's
orbital elements, and P (Di) is the occurrence probability
of our computed lifetimes. Of course, we have N TNOs
to consider in this analysis, so what we really need to
compute is the posterior probability distribution as it
depends on the numerically estimated lifetimes of all N
objects:
P (AD1, D2...DN ) =
P (D1, D2...DNA)P (A)
P (D1, D2...DN )
,
P (ATi, I) =
P (Ti, IA)P (A)
P (Ti, I)
(1)
which, using the definition of Bayes' theorem, reduces
Planet Nine Stability Survey
7
Outcomes of ETNOs in presence of Neptune and Planet Nine
TABLE 2
Object
Migration (a > 100 AU) Close Encounter Collision with Central Body Ejection
Survive to 4.5 Gyr
2003 VB12 (Sedna)
2004 VN112
2007 TG422
2010 GB174
2012 VP113
2013 RF98
2013 FT28
2014 SR349
16.0%
14.9%
29.3%
17.8%
7.1%
14.0%
17.2%
22.0%
19.3%
30.6%
32.9%
31.2%
21.1%
40.6%
30.7%
39.5%
0.1%
0.7%
0.0%
0.4%
0.5%
0.0%
0.0%
0.7%
19.3%
25.4%
23.1%
32.7%
14.8%
26.7%
32.2%
19.8%
45.3%
28.4%
14.6%
17.9%
56.5%
18.7%
19.9%
18.1%
Note. -- For each TNO used in the dynamical simulations, the percentage of integrations that ended in each major instability
outcome.
into the form:
P (AD1, D2...DN ) =
P (D1) × P (D2) × ...P (DN )
×
P (D1A) × P (D2A) × ...P (DNA)
P (D1, D2...DN )
P (A)N−1
(3)
We can say that D1, D2...DN are conditionally inde-
pendent, since we treated all TNOs as test particles
in the simulations used to generate {Di}. Since each
TNO has zero mass in the simulations, removing one
will not alter the lifetime maps for the other objects.
As a result, the first term of the right-hand side of
Equation 3 can be treated as a normalization coefficient,
which is needed only when comparing different versions
of P (AD1, D2...DN ) constructed with different numbers
N of TNOs. We need only a relative posterior probability
distribution, which will identify the most likely realiza-
tion of Planet Nine and identify the parameter space in
which it is likely to reside.
(cid:81)n=N
The final result we need to compute, then, is:
n=1 P (DiA)
P (A)N−1
P (AD1, D2...DN ) = a
(4)
when D denotes the lifetime of a TNO in the presence of
Planet Nine as measured from simulations, i denotes the
TNO considered, A denotes the orbital elements (a, e) of
Planet Nine. P (A) is a known quantity, and P (DiA) can
be measured from the stability maps constructed using
the N-body integrations.
3.2. Converting TNO lifetime maps to probability
distributions
Equation 4, which can be used to construct our goal
posterior probability distribution, requires P (DiA) for
each TNO in the sample. P (DiA) is the probability of
the TNOs' lifetimes as they depend on the orbital ele-
ments of Planet Nine. To compute this term, we must
convert the lifetime maps presented in Figure 2 into prob-
ability distributions.
Each realization of Planet Nine will lead to a differ-
ent time of dynamical instability (defined as the time at
which an object experiences one of the four instability
mechanisms described in Section 2.2), with some objects
never experiencing dynamical instability. In Figure 3, we
plot a histogram of the object lifetimes for 2013 RF98,
as derived from the 1500 numerical integrations run with
varying realizations of Planet Nine.
Fig. 3. -- A histogram of lifetimes (measured from time t = 0
at the start of the N-body simulation) for 2013 RF98. Barring a
pile-up at lifetimes of t = 4.5 Gyr (corresponding to realizations
that were stable for the duration of the simulation; these were not
included in the fit), the decay appears to have an approximately
exponential trend.
Although the longer object lifetimes can intuitively be
interpreted as corresponding to more likely realizations
of Planet Nine, we would like to convert these lifetimes
into a relative probability function. Since an exponential
decay trend appears to be a good, empirical fit for the
lifetimes of an object marginalized over all integrations
(regardless of the fact that the realization of Planet Nine
varies between the different trials), we can fit a decay
constant λ from the lifetime histogram for all decaying
realizations of a single TNO:
fdecay(t) = C0e−λt
(5)
when fdecay(t) is the fraction of realizations decaying at
each time t, C0 a normalization constant, and λ the de-
cay constant to be fit. For each object in our sample, the
histogram of instability times was fit with exponential
curve using a simple Levenburg-Marquardt optimization
scheme, with all objects that do not experience dynam-
ical instability (which have lifetimes t = 4.5 Gyr) ex-
cluded from the fit. The results of this fit for 2013 RF98
are plotted in Figure 3. For this particular object, if
we extend the exponential curve to infinite time, we ex-
pect only 0.4% of these dynamically stable objects to
become dynamically unstable, indicating that the ma-
jority of objects that have not decayed after 4.5 Gyr are
truly dynamically stable.
8
Becker et al.
(cid:26)Ns/N
To construct a final probability distribution, we need
to account for the fact that for each object, some fraction
of trials were dynamically stable for the entire 4.5 Gyr
integration length. For this reason, we construct a piece-
wise probability function for P (DiA), the probability we
get these lifetimes from our simulation given a particular
set of A, Planet Nine's orbital elements:
P (DiA) =
if Di=4.5 Gyr
else
C1 × e−λDi
(6)
when P (DiA) is the probability of computing the ob-
ject lifetime Di for the ith TNO in the presence of a
Planet Nine with orbital elements A, Ns the number of
integrations for which the object did not experience dy-
namical instability, N the total number of integrations,
λ the decay constant determined from the previous fit,
and C1 is a normalization constant of the form:
(cid:34)(cid:90) 4.5 Gyr
(cid:35)−1
e−λtdt
(7)
C1 =
N − Ns
N
×
0
The substitution of Equation 6 into Equation 4 results
in a final expression of the posterior probability distribu-
tion P (AD1, D2...DN ). For the eight objects considered
in this work, this final distribution is plotted in Figure
4. It is important to note that this process required mul-
tiple steps of normalization, which depend on P (A), the
priors, and the populations of TNOs used to derive con-
straints. Since we (by necessity) treated the objects used
in this analysis as conditionally independent, the poste-
rior presented in Figure 4 provides relative, rather than
absolute, measures of the likelihood of each (a, e) real-
ization of Planet Nine. The derived decay constant must
be re-derived if different populations of TNOs or differ-
ent orbits of Planet Nine are tested in the future; the
values used in this work are particular to our sample of
eight TNOs in the presence of our particular population
of Planet Nines. For this reason, we do not provide the
decay constants in this work, as they cannot be used
for these objects in general but only for our particular
choices of Planet Nine's priors.
3.3. The posterior probability distribution for Planet
Nine's orbital elements
Figure 4 presents the final posterior probability dis-
tribution for Planet Nine's orbital elements (a, e), based
on the observed dynamical stability of the eight TNOs
considered in this work. This distribution provides a
relative measure of the likelihood of differing combina-
tions of Planet Nine's semi-major axis and eccentricity.
The nominal orbit of Planet Nine (700 AU, 0.6 eccen-
tricity) appears to lie in a less preferred region, with
slightly smaller eccentricities (0.3-0.4) being preferred.
Remarkably, the large-a, low-e orbits appear to be ex-
cluded based on our dynamical stability arguments. This
is roughly consistent with the posterior generated us-
ing clustering arguments, provided in Brown & Batygin
(2016), which also preferred a middling eccentricity and
excluded large-a, low-e iterations of Planet Nine.
3.4. Comparing our prediction of Planet Nine's Orbit
with constraints derived via different methods
Fig. 4. -- The overall stability posterior for the semi-major axis
and eccentricity of Planet Nine. This posterior was constructed by
taking a summation of the posteriors for each individual object,
including the six objects used in Batygin & Brown (2016a) and the
two new high-a, low-e objects from Sheppard & Trujillo (2016).
The posterior probability distribution we present in
Figure 4 was generated using dynamical stability argu-
ments. Specifically, Planet Nine realizations that allow
the observed TNOs to remain dynamically stable are con-
sidered to be more likely than those that cause dynami-
cal instabilities. Brown & Batygin (2016) also provided
a probability map showing the most likely regions for
Planet Nine based on the observed orbital alignment of
the TNOs. In Figure 5, we compare the two probability
distributions presented in Brown & Batygin (2016) to the
distribution derived in this work (along with additional
contours described below). For these distributions, we
plot a single contour in Figure 5 using the value 1σ be-
low the maximum probability of the distribution. The
level of this contour is chosen so that the single contours
visually reflect the highest probability regions for each
posterior type. The choice of the (maximum - 1σ) level
is representative, but by no means the only way to visu-
alize the comparisons between the posterior types.
The regions showing the most overlap between the 10
M⊕ alignment result from Brown & Batygin (2016) and
our dynamical stability result are those with a semi-
major axis between 525 -- 675 AU and an eccentricity of
0.40 -- 0.55, and the region with a semi-major axis be-
tween 700 -- 800 AU and an eccentricity of 0.25 -- 0.35. A
recent paper by Millholland & Laughlin (2017) also used
dynamical stability arguments and resonance considera-
tions to choose a best-fit (a, e) of (654 AU, 0.45). This
point is also plotted in Figure 5 and appears to be con-
sistent with the region showing overlap between both
the dynamical stability results of this work and those
of (Brown & Batygin 2016).
The posteriors plotted in Figure 5 include four curves
based on N-body simulations and one analytical curve.
The analytical curve encompasses the regime where the
equation of motion dω/dt ≈ 0. It is important to note
that this analytic approximation is not a true analysis of
the dynamics of the system, as the derivation (the com-
plete form of which is presented in Appendix A) assumes
all bodies are coplanar, and solves for alignment rather
Planet Nine Stability Survey
than anti-alignment. Thus, this analytic model is more
an order-of-magnitude estimate of the effect we expect
to see, rather than a true prediction.
The functional form for dω/dt is derived from the sec-
ular Hamiltonian, and assumes the form:
mia2
4(cid:88)
i (e2 − 1)−2a−2
(cid:19)4 √
(cid:18) a
(cid:19)3 √
1 − e2
9)5/2
a9
i=1
(1 − e2
1 − e2
9)3/2
,
(1 − e2
a9
dω
dt
∝ 3
4Mc
m9
Mc
e9
e
(cid:18) a
− 15
16
− 3
4
m9
Mc
(1 − 9e2/4) cos(ω − ω9)
9
about 2 × 10−5 (for alignment among all eight TNOs).
As a result, the numerical simulations show an overabun-
dance of alignment in certain, preferred regions (denoted
by the green contour of Figure 5).
The constraints provided by the three methods in this
work and the two literature results show rough agree-
ment: the region of Planet Nine parameter space look-
ing most attractive extends over the range 500 -- 700 AU
in semi-major axis, and 0.3-0.6 in eccentricity. Notably,
low eccentricities are disfavored. Notice also that the
contours preferred for TNO stability are somewhat par-
allel to those for orbital alignment. In order for Planet
Nine to be close enough to the TNOs to enforce align-
ment, it must be close enough that the TNO orbit is close
to instability. As a result, the preferred region is given
by the boundary between the stablity contours and the
alignment contours in Figure 5.
4. END STATES FOR TRANS-NEPTUNIAN OBJECTS
UNDER THE INFLUENCE OF PLANET NINE
The numerical simulations that we use to create distri-
butions of TNO lifetime in the presence of Planet Nine
evaluate the long-term evolution of these TNOs. As dis-
cussed in Section 2, there are five main outcomes for
these TNOs: (1) migration in semi-major axis, which is
defined in the simulations as the orbit's semi-major axis
attaining a value more than 100 AU from the starting
orbit; (2) a close encounter (defined as passing within 3
Hill radii of the larger body) with Planet Nine or Nep-
tune, which would result in the TNO being captured by
or colliding with the large planet; (3) collision with the
central body; (4) ejection from the solar system, where
the ejection radius is taken to be 10,000 AU3; and fi-
nally, (5) dynamical stability, where none of the afore-
mentioned effects occur over the 4.5 Gyr timescale of the
simulation. This final result (5) corresponds to scenarios
that are consistent with observations, where the TNO
can remain in its orbit over solar system lifetimes. This
scenario allows the secular alignment of the TNOs' lon-
gitude of perihelion by Planet Nine. However, if this last
result does not occur, then the TNO will experience one
of the first four (1-4) outcomes.
Within these four dynamical
instability outcomes,
there is further stratification in the effects of each mecha-
nism. The violent ends (collision with a planet, the star,
or ejection from the system) generally remove the TNO
from the solar system entirely, while the migratory out-
come is more nuanced. TNOs that are in the process
of migrating cannot necessarily be used as evidence of
the Planet Nine hypothesis, since their orbital elements
are in flux and may not have had sufficient time in their
current orbits to attain the alignment that serves as a
hallmark of Planet Nine's influence.
2007 TG422 was one of the more interesting objects in
the sample of TNOs considered in this work. Although
found in Sheppard & Trujillo (2016) to be dynamically
unstable in the presence of Neptune, we reevaluated the
stability of 2007 TG422 in the presence of Planet Nine
in Section 2, and found that adding Planet Nine to the
3 This choice in ejection radius will remove objects which become
unbound; it is important to note that there may be second order ef-
fects due to stellar encounters (see, for example, Li & Adams 2016).
The effect on TNO motion due to these external perturbations is
expected to be small, and is neglected in this work.
(8)
where the subscript 9 denotes the orbital elements of
Planet Nine, the subscript i within the summation in the
first term denotes the four giant planets, c denotes the
central body, and a lack of subscript denotes the TNO for
which the equation of motion is written. This equation
of motion can be split into two parts: dωSS/dt, which
denotes precession due to the effect of the gas giants, and
dω9/dt, which denotes the precession due to the effect
of Planet Nine. The magnitudes of these terms can be
written in the form
4(cid:88)
i=1
m9
Mc
dωSS
dt
dω9
dt
∝ 3
4Mc
∝ 15e9
16e
+
3
4
m9
Mc
mia2
(cid:18) a
(cid:18) a
a9
i (e2 − 1)−2a−2
(cid:19)4 √
1 − e2
(cid:19)3 √
(1 − e2
9)5/2
1 − e2
9)3/2
(1 − e2
a9
(1 − 9e2
4
) cos(ω − ω9)
(9)
where dωT N O/dt = dωSS/dt − dω9/dt. At the point
where dωSS/dt ≈ dω9/dt, the precession rates due to
Planet Nine and the inner solar system cancel each other
out, and the TNO orbit is not expected to precess. For
each TNO, we can construct a curve in the (a, e) plane
for which these precession rates cancel. The region pre-
sented in Figure 5 is the superposition of these curves for
all eight TNOs, and it encompasses the range of Planet
Nine realizations which will allow alignment by prevent-
ing precession.
Figure 5 also presents (in green) a posterior based on
alignment of the TNOs, as derived from our set of nu-
merical simulations. Our simulations were intended to
test dynamical stability, but one output of the N-body
simulation is the orbital elements ω and Ω of the TNOs
over time. Using a similar technique to that used to
construct the dynamical stability posterior in Figure 4,
we tested how well aligned the TNOs were for each re-
alization of Planet Nine. Alignment was measured by
looking at the fraction of time for which all eight TNOs
were aligned with each other (and thus anti-aligned with
Planet Nine: we counted dispersions of less than 94 de-
grees in = ω + Ω between all eight TNOs as aligned).
We constructed a contour plot of the percentage of the
integrations during which alignment was visible for all
eight TNOs, which exhibited between near 0% alignment
to a maximum of ∼10% alignment in the green regions.
The alignment rate expected for pure chance would be
10
Becker et al.
Fig. 5. -- A comparison between the preferred regions for Planet Nine's orbit, as computed using a variety of different methods. Bold
In red, we plot the region that maximizes the survival probability for the
labels indicate that the posterior was derived in this work.
TNOs in our N-body simulations (based on their dynamical stability; see Figure 4).
In green, we plot the realizations of Planet Nine
orbital parameters in our N-body simulations that allowed the TNOs to be aligned as observed in nature. In black, we plot the analytic
approximation for the region where we expect the observed alignment to occur (see Equation 8 and its derivation in the Appendix B). In
blue, we plot the region reported in Brown & Batygin (2016) that allows the alignment of TNOs in a suite of numerical N-body simulations
(unlike the green curve, these simulations used randomized test particles). The purple point denotes the best-fit orbit as reported by
Millholland & Laughlin (2017), which was found by optimizing the resonant behavior of the TNOs in the presence of Planet Nine. Each
method prefers somewhat different regimes of parameter space, with the dynamical stability argument allowing the largest region. There
is significant overlap between the results derived from the different methods. The optimal orbital elements for Planet Nine exist in the
overlap region, which corresponds to eccentricity in the range e9 = 0.4 -- 0.6 and semimajor axis in the range a9 = 500 -- 700 AU.
solar system can actually stabilize the orbit of this ob-
ject. 2007 TG422 has a particularly high semi-major axis
and eccentricity, suggesting that its dynamical instabil-
ity might be due to orbit crossing with other planets.
However, this is not the case. The most common dy-
namical instability mechanism for 2007 TG422 is actually
migration, leading this object to remain present in the
solar system but wander from its starting orbit. This
migratory outcome is likely what Sheppard & Trujillo
(2016) found in their work, and we reproduce this insta-
bility in the case of large-a, low-e realizations of Planet
Nine (which is dynamically very similar to the case of
no Planet Nine, as considered in Sheppard & Trujillo
(2016)).
The top panels of Figure 6 show the instability life-
time map for 2007 TG422 when only violent instability
methods are considered (top panel) and when migratory
instabilities are also considered (middle panel). The sig-
nificantly shortened lifetimes present in this middle panel
show that migration is the main explanation for 2007
TG422's dynamical instability in the presence of Planet
Nine. The bottom panel of Figure 6 shows the difference
in expected dynamical stability lifetime for this object
between the cases depicted in the top two panels: when
this difference is low, migration does not have a large
effect in ending the integrations, and when it is high,
migration is a very common outcome for 2007 TG422 in
the presence of that particular Planet Nine realization.
This difference plot highlights the regions where migra-
tion acts as the main cause of dynamical instability. The
large-a, low-e realizations of Planet Nine tend to cause
migration of 2007 TG422, but not violent dynamical in-
stabilities.
The susceptibility of an object to migration in the pres-
ence of Planet Nine can also be summarized as ¯δt, the
difference in median lifetime (over all Planet Nine re-
alizations) for an object between the case when migra-
tion is considered to be a dynamically unstable outcome
and when it is ignored. Larger values of ¯δt indicate that
the TNO is more susceptible to significant migration in
semi-major axis (δa > 100 AU), and that such migra-
tions significantly change the dynamical stability map of
that object. The choice of δa > 100 AU as the threshold
criterion for migration is somewhat arbitrary. It is cho-
sen to represent the condition for which an object has
drifted significantly from its observed orbit. Migration
Planet Nine Stability Survey
11
in semi-major axis with deviations greater than 100 AU
generally results in an orbit with significantly different
orbital elements, rather than oscillation of the orbital el-
ements around well-defined central values. As a result,
objects that migrate by more than 100 AU are unlikely
to remain part of the same dynamical class of object.
For 2007 TG422 , ¯δt = 1.07 Gyr, meaning that Planet
Nine incites migration in this object often enough to de-
crease its dynamically stable lifetime by more than one
In contrast, 2012 VP112 has a ¯δt = 32
billion years.
Myr, indicating that migration does not play a large role
in its outcomes. The values of ¯δt for each object in our
sample are reported in Table 1, which presents in its last
two columns both ¯δt and the percentage of trials out of
all our simulations that are dynamically stable. As the
percentage increases (or, as an object is more stable in
the presence of any considered Planet Nine realization),
the value of ¯δt tends to decrease (or, the TNO experi-
ences migration as an instability outcome with a lower
frequency). This suggests that susceptibility to migra-
tion is a major factor leading to differences in orbit life-
times between different TNOs in the presence of Planet
Nine.
In this work, we have made the assumption that since
the eight TNOs in our sample exhibit alignment in lon-
gitude of perihelion attributable to Planet Nine, these
objects by necessity have lived in their current orbits
for a significant length of time. In constructing the sta-
bility posterior presented in Figure 4, we assumed that
since significant, unbounded migration alters orbits and
potentially disrupts this alignment, a migratory insta-
bility provides equal information against Planet Nine's
orbital elements as does a violent dynamical instability.
In this way, we used migratory instabilities (such as those
caused by large-a, low-e realizations of Planet Nine for
2007 TG422) as evidence against the iterations of Planet
Nine that excited the migration.
The TNOs affected by Planet Nine appear to fall into
two categories: some of them (Sedna, 2012 VP113) are
generally stable against this migratory process, while
others (2007 TG422, 2013 RF98) can be easily caused
to migrate in semi-major axis with an improperly cho-
sen realization of Planet Nine. The best orbit of Planet
Nine can be determined not only from the orbit-crossing
constraint that rules out small-a, large-e orbits for all
TNOs, but also by the constraint from this second pop-
ulation of objects, which are unstable in the presence of
large-a, low-e Planet Nine orbits (and are additionally
unstable in the presence of no Planet Nine at all). These
two populations explain the unintuitive structure of the
posterior probability distribution given in Figure 4.
Even though we require a Planet Nine iteration that
does not cause the aligned TNOs to migrate significantly
(more than 100 AU) over secular timescales, we expect
there is a further population of objects that do migrate
in the presence of Planet Nine.
Indeed, some popula-
tions of objects in our solar system can be explained by
this process. Batygin & Brown (2016b) uses the nomi-
nal Planet Nine orbit from Batygin & Brown (2016a) to
explain the existence of a population of highly inclined
KBOs with a ≤100: these objects can be explained as a
migratory end-state of objects that started as members
of the extreme TNO populations. The fact that some
TNOs appear to experience migratory instabilities un-
der the influence of Planet Nine is not inconsistent with
our assumptions, but it does engender further questions.
Clearly, some degree of migration is acceptable and will
not alter the orbital alignment of the TNOs - the 100 AU
threshold we use allows significant movement in semi-
major axis without declaring objects dynamically unsta-
ble. However, the fact that objects can migrate in semi-
major axis but remain confined to a comparably narrow
range (disallowing above 100 AU of migration does not
allow objects to move into entirely different object pop-
ulations as seen in Batygin & Brown 2016b) begs the
question: how are objects moving when they migrate in
what we have defined as a dynamically stable way? The
answer to this question will be addressed in the following
section.
5. PROXIMITY OF THE TNO ORBITS TO RESONANCES
WITH PLANET NINE
In the previous section, we identified that for some
TNOs, migration in semi-major axis is an outcome for
a significant fraction of integrations.
In this section,
we delve deeper into the nature of that migration and
consider the question of resonance between the TNOs
and Planet Nine. Specifically, if the TNOs migrate in
semi-major axis, they are likely to pass through the lo-
cations of mean-motion commensurabilities with Planet
Nine, and one might expect the TNOs to fall into the
stable configurations afforded by resonances, and stop
migrating.
The fact that proximity to resonance boosts dynami-
cal stability can be directly applied to the Planet Nine -
TNO system. In particular, Malhotra et al. (2016) con-
sidered the possibility that Planet Nine should be at an
orbital radius that would allow it to be closest to low-
order resonances for several TNOs, based on the cur-
rently measured orbits of those TNOs. The benefit to
this configuration is that Planet Nine could stabilize the
orbits of TNOs such as 2007 TG422, which might other-
wise migrate in semi-major axis. Similarly, Millholland
& Laughlin (2017) used numerical N-body simulations
to determine the best orbit of Planet Nine by testing
which locations close to resonances allow the observed
alignment of the TNOs. Millholland & Laughlin (2017)
found the best semi-major axis of Planet Nine to be a9 ∼
654 AU (with a e9 ∼ 0.45), and Malhotra et al. (2016)
found the best-fit orbit to have a semi-major axis of a9 ∼
665 AU.
Clearly, resonance is an important aspect of the Planet
Nine problem, as suggested by Batygin & Brown (2016a),
Beust (2016), Malhotra et al. (2016), and Millholland &
Laughlin (2017).
In Figure 7, we plot two histograms
of the orbital period ratio (P9/PT N O) for each of the
eight TNOs in our sample.
In the top panel, we plot
random draws from our initial conditions that were used
to initialize the N-body simulations described in Section
2. The total number of draws in the top histogram was
chosen to match the number of dynamically stable time-
steps for each individual TNO, but the draws are from a
raw distribution of the ratio P9/PT N O for all numerical
trials, including those that are not dynamically stable.
This top panel does not include the effect of certain ra-
tios becoming more common due to gravitational inter-
actions. In the bottom panel, we plot the dynamically
12
Becker et al.
stable trials for each TNO, with the period ratio sam-
pled every million years. The sharp peaks in this second
histogram demonstrate clear overabundances of partic-
ular period ratios P9/PT N O. These values occur near
resonances -- in the trials where dynamical stability is
found -- and is markedly different from the continuum
of period ratios shown in the upper panel. Keep in mind
that the initial conditions of our simulations were chosen
independently of resonance locations, unlike the set of
simulations in Millholland & Laughlin (2017). As a re-
sult, the behavior shown in the figure indicates that sys-
tems found to be dynamically stable in the simulations
also show a clear preference for near-resonant locations.
The results of this work show an important departure
from the assumptions used previously (Malhotra et al.
2016; Millholland & Laughlin 2017). This earlier work
assumes that the TNOs remain in a single resonant con-
figuration over the lifetime of the Solar System. For ex-
ample, Sedna might live in either the 9:8 or 6:5 mean
motion commensurability, depending on the semi-major
axis of Planet Nine, but it is considered to remain in
a single resonance. In this approximation, the orbits of
the extreme TNOs that we observe today are consistent
with their past orbits. This assumption allows for the
estimate found in Malhotra et al. (2016) to be carried
out: if the TNOs reside in the same orbits over their en-
tire lifetimes, and if they must be near resonance, then
their current orbital properties can be used to compute
the expected orbital elements of Planet Nine. However,
our simulations show that TNOs do not always remain in
the same orbits: Although they often remain near some
resonance with Planet Nine, the TNOs change orbits and
hence change resonances.
In Figure 8, we plot time series for 2004 VN112 (top),
2007 TG422 (middle), and Sedna (bottom) to demon-
strate the potential resonant outcomes. The behavior
of 2004 VN112 is consistent with the assumption made
by Malhotra et al. (2016) and Millholland & Laughlin
(2017): that TNOs would live in a single resonance for
the age of the solar system. The other two TNOs plot-
ted in Figure 8 show behavior we call 'resonance hop-
ping,' where a TNO attains multiple mean-motion com-
mensurabilities over the course of the simulation. When
liberated from one mean-motion commensurability, both
TNOs are captured into another resonance instead of be-
ing ejected from the system entirely. Planet Nine's semi-
major axis did not change over the course of a single
simulation, so all the change in P9/PT N O within a single
integration is due to migration by the TNO.
As illustrated in Figure 2, each TNO has a different
dynamically stability map, preferring different regions of
Planet Nine's potential parameter space. Similarly, when
we take the subset of dynamically stable integrations for
each TNO, each TNO has a different behavior relating
to resonance. In this work, we take the definition of be-
ing 'in' resonance to be living close to a resonant period
ratio; as with the exoplanetary systems near resonance,
the boost in dynamically stability provided by proximity
to resonance applies even when systems are not in a per-
fect resonance. Malhotra et al. (2016) and Millholland
& Laughlin (2017) use a criterion for the proximity to
Fig. 6. -- The lifetime of 2007 TG422 in the presence of realiza-
tions of Planet Nine with semi major axis between 450 and 1200
AU, and eccentricities between 0 and 1. (top panel) Stability life-
times when dynamical instability is only caused by violent ends for
2007 TG422, including collision with a solar system planet, collision
with the sun, or ejection from the solar system (the outer bound-
ary of which is defined to be 10000 AU). (middle panel) Stability
lifetimes, when dynamical instability is caused by the violent end
depicted in the top panel and also migration in semi-major axis by
more than 100 AU. (bottom panel) The difference δt in stability
lifetimes between the two cases, demonstrating which realizations
of Planet Nine are most likely to cause 2007 TG422 to change its
orbit significantly. Planet Nine realizations with low eccentricity
and large semi-major axis cause 2007 TG422's orbit to migrate, but
not meet with a violent end.
2007 TG422 Planet Nine Stability Survey
13
Fig. 7. -- (top panel) A sample of period ratios P9/PT N O, drawn from the initial conditions of the N-body simulations. This histogram
shows the period ratio distribution that we would expect to see from the simulations if no period ratio were more dynamically stable than
any other period ratio. This set serves as the control, and does not have the condition of dynamical stability imposed. (bottom panel)
Period ratios P9/PT N O for the dynamically stable integrations at time-steps of one million years, demonstrating a peaked distribution, as
some period ratios are preferred to others. In both panels, colors correspond to the values for each TNO, and the area of histograms was
normalized.
resonance that is close enough to afford such benefits:
∆ares ≈ 0.007aTNO m9aTNOA
3M(cid:12)a9
1/2
(10)
when ∆ares is the width, in AU, of the band of space close
enough to a resonance to count as being 'near' said res-
onance, subscript T N O denotes the TNO's semi-major
axis a and the subscript 9 denotes Planet Nine's semi-
major axis a9 and mass m9. A is a unitless coefficient.
We choose to use A = 3 as done in Malhotra et al. (2016)
and Millholland & Laughlin (2017). The numerical value
of ∆ares tends to be close to 5-10 AU for the TNOs in
our sample, meaning that the simulation results allow us
to determine the nearest resonance and identify TNOs
that 'hop' between resonances. Figure 8 shows three ex-
amples from our set of simulations: 2004 VN112 does not
change resonances during the integration, 2007 TG422
changes resonances three times, and Sedna changes res-
onance once. For two objects in our sample (2004 VN112
and 2012 VP113), we see no hopping behavior.
However, it is important to note that the simulations
used to construct Figure 7 and the curves in Figure 8
are the set run for this paper, which uses the quadrupole
moment of the central part of the system (J2) to replace
the active motions of Jupiter, Saturn and Uranus (JSU).
This approximation allows the simulations in this work
to be completed in a total of roughly 100,000 total CPU
hours, rather than the nearly 2 million CPU hours that
would be required to run the full integrations with all ac-
tive giant planets. This approximation is appropriate for
the tests of dynamical stability and alignment considered
thus far in this work, but when considering the question
of resonance, some discrepancies arise. In Figure 9, we
present a comparison between a few test simulations run
with the J2 approximation and the JSU case of active
particles for Jupiter, Saturn, and Uranus. The test sim-
ulations were run for 1 Gyr each, with a single realization
of Planet Nine (a = 700 AU, e = 0.5) and otherwise iden-
tical to the cases run in the previous set of simulations.
For the JSU set of simulations, we lowered the time-step
to 20 days.
When active JSU particles are included in the integra-
tions, two major differences occur as compared to the J2
approximation: (1) the period ratios are not as tightly
024681012Period Ratio (P9/PTNO)02000400060008000100001200014000FrequencyInitial conditionsFT28GB174RF98SEDNASR349TG422VN112VP113123456Period Ratio (P9/PTNO)02000400060008000100001200014000Frequency311131147217452175225545141125236Dynamically stableFT28GB174RF98SEDNASR349TG422VN112VP11314
Becker et al.
Fig. 8. -- For three integrations (top curve: 2004 VN112; middle
curve: 2007 TG422; bottom curve: Sedna), we plot the period ratio
P9/PT N O as function of simulation time. The numbers denote
resonances within ∆ares of the semi-major axis of each TNO, at
each time-step. The top time series, showing the evolution of 2004
VN112 during a typical integration, demonstrates how an object
might remain in a single mean-motion commensurability for the
entire lifetime of the solar system. The middle series, for 2007
TG422, shows the behavior we call 'resonance hopping,' where a
TNO attains multiple mean-motion commensurabilities over the
course of the simulation. The bottom series shows a less extreme
version of this 'resonance hopping,' where Sedna switches between
two commensurabilities during the solar system lifetime.
In all
cases, the period of Planet Nine does not change over the course of
the simulation: it is the motion of the TNO that leads to changing
values of P9/PT N O.
confined, experiencing a larger degree of scatter even
while living in a single apparent resonance; (2) the num-
ber of times objects hop between resonances can be both
increased (due to repeated accelerations from Uranus)
and decreased (as shown in the top panel of Figure 9,
hops we resolve in the J2 case are un-physical in the JSU
case).
Regarding point (1), the second panel of Figure 9 il-
lustrates a case where in both the J2 and JSU cases, a
TNO lives close to the 8/5 mean motion resonance. In
the J2 case, the TNO stays within 0.3% of its average
period ratio. In the JSU case, the TNO stays within 5%
of its averege period ratio (this value is larger than in
the J2 case due to the inclusion of accelerations from the
inner three giant planets). Due to this complication, we
have chosen the bin size in Figure 7 to be commensurate
with the period ratio confinement experienced by TNOs
in the JSU case, which is typically around 8%.
Regarding point (2), it is unclear without running a
large suite of JSU integrations how the hopping fre-
quency changes with the inclusion of the giant planets.
For this reason, we do not present in this work detailed
results of the resonance hopping in our simulations. In
order to accurately assess this behavior, a complete set
of simulations with active giants planets (JSU) should be
carried out.
Our simulations show that dynamically stable integra-
tions of the TNOs tend to attain mean-motion commen-
surabilities. Depending on the TNO, it may attain a
single resonant location and stay there, or it may hop
between resonant locations. This 'resonance hopping' is
an important effect, and for TNOs that exhibit this be-
havior, their past semi-major axis may be different than
the current values. The numerical computation of the
Fig. 9. -- A check of our integrations (which used solar J2 in place
of the inner three giant planets) and integrations run with active
Jupiter, Saturn, and Uranus (JSU) show that although both sets
exhibit the same hopping behavior, the J2 approximation under-
estimates the noise in period ratio and overestimates the degree to
which the resonances can be differentiated.
specifics of this behavior should be a fruitful avenue for
future work.
6. CONCLUSIONS
In this work, we have evaluated the dynamical sta-
bility and orbital alignment of eight TNOs (Sedna, 2004
VN112, 2007 TG422, 2010 GB174, 2012 VP113, 2013 RF98,
2013 FT28, and 2014 SR349) in the presence of a Monte
Carlo assortment of Planet Nine realizations with vary-
ing semi-major axis and eccentricity. We used the results
to predict the most probable (a, e) of Planet Nine by de-
riving the posterior probability distributions for Planet
Nine's orbital elements (a, e). The distribution based on
dynamical stability considerations for the TNOs is pre-
sented in Figure 4. We have also constructed an anal-
ogous probability distribution based on the requirement
that the orbits of the TNOs remain aligned. Both of
these posterior distributions demonstrate that the pre-
ferred orbits for Planet Nine have intermediate values
of eccentricity (0.3 < e < 0.5) and semi-major axis
(650 < a < 900 AU), as shown in Figure 5. Moreover,
0.00.51.01.52.02.53.03.54.04.5Time (Gyr)1.01.52.02.53.03.5Period Ratio (P9/PTNO)2004 VN1123/12007 TG42217/87/311/52/1Sedna9/86/50.00.20.40.60.81.0Time (Gyr)3.03.13.23.33.43.53.63.73.83.9Period ratio (P9/PTNO)2013 FT28J-S-U active planetsJ2 approximation0.00.20.40.60.81.0Time (Gyr)1.351.401.451.501.551.601.651.701.751.80Period ratio (P9/PTNO)SednaJ-S-U active planetsJ2 approximation0.00.20.40.60.81.0Time (Gyr)1.52.02.53.03.54.0Period ratio (P9/PTNO)2010 GB174J-S-U active planetsJ2 approximationPlanet Nine Stability Survey
15
these values are roughly consistent with the regime sug-
gested in Brown & Batygin (2016), which constructed
its probability map using clustering arguments only. Our
stability posterior and that from Brown & Batygin (2016)
were constructed based on different fundamental orbital
properties (dynamical stability and secular evolution pat-
terns, respectively). Despite this significant difference in
construction, the two results are consistent, in that they
both prefer non-zero eccentricities and a similar range
in semi-major axis for Planet Nine. The comparison be-
tween our results and those of Brown & Batygin (2016)
is shown in Figure 5. Notably, similar dynamical stabil-
ity arguments in Millholland & Laughlin (2017) produce
a best-fit Planet Nine of 654 AU and 0.45 eccentricity,
which is consistent with the overlap region between the
results of this work and of Brown & Batygin (2016).
Using numerical N-body simulations, we also demon-
strated that 2007 TG422 and 2013 RF98, while found in
the past to be dynamically unstable in the presence of
Neptune alone, can attain dynamically stable states in
the presence of Planet Nine. Our simulation results sup-
port the prediction of Sheppard & Trujillo (2016) that
since 2007 TG422 and 2013 RF98 exhibit the same orbital
clustering as the dynamically stable TNOs, Planet Nine
likely dominates over Neptune interactions. In addition,
we find that different TNOs exhibit very different sta-
bility maps, with some objects (such as Sedna and 2012
VP113) contributing relatively little unique information
to the stability posterior and others (such as 2007 TG422
and 2013 RF98) exhibiting unintuitive preference against
large-a, low-e orbits of Planet Nine. These two categories
suggest that there may be two dynamical classes of ob-
jects in this TNO sample, which interact differently with
Planet Nine. However, we have considered in this work
only a small (N = 8) number of objects that fit into
the desired high-a, trans-Neptunian-q, apsidally aligned
category identified in Batygin & Brown (2016a). The dis-
covery of additional objects in this population (expected
in the near future) will allow for a more robust test of
this two-population hypothesis.
We have also evaluated the different dynamical out-
comes for these extreme TNOs in the presence of Planet
Nine. The objects that are dynamically unstable in the
presence of large-a, low-e orbits of Planet Nine (2007
TG422 and 2013 RF98) tend to experience migration
rather than violent collisions or ejections as their main
outcome in dynamically unstable cases. These objects
are also dynamically unstable in the presence of only
Neptune and the other giant planets, i.e., in the absence
of Planet Nine. In Table 1, we present the difference ¯δt
in average dynamical lifetime between the case where mi-
gration is considered to be a dynamical instability mech-
anism and when it is not. Table 2 presents the relative
occurrence rates for each type of outcome. For cases
where the TNOs are not stable over the lifetime of the
Solar System, the fraction of trials that lose objects to
migration (with a > 100 AU), close encounters with giant
planets, and ejection from the system are roughly com-
parable. A small minority of the simulations end with
accretion onto the Sun (less than 1%).
Next, we suggest a generalized description for the inter-
actions between the TNOs and Planet Nine. We propose
that the paradigm is neither that (1) mean motion reso-
nance is unimportant, nor (2) TNOs reside in a single res-
onance with Planet Nine for the age of the solar system.
Instead, while some TNOs (such as 2004 VN112 and 2012
VP113) can sometimes live in a single resonance for solar
system lifetimes, others exhibit a behavior that we call
'resonance hopping.' This term means that the TNO is
near-continually in close proximity to a mean motion res-
onance (Figure 7), but it is not necessarily near the same
resonance for the age of the Solar System (Figure 8). In-
stead, the TNOs can transition between closely-spaced
resonances, often those described by relatively large in-
teger ratios. The long-term effect of this process is that
the orbital anti-alignment caused by Planet Nine is able
to persist, but the TNO is protected against small kicks
in energy provided by interactions with Neptune. In this
paradigm, an interaction with Neptune might lead to the
movement of a TNO into a new resonance, but not to its
ejection from the solar system (see also Malhotra et al.
2016; Millholland & Laughlin 2017). A useful avenue for
future work would be the full numerical computation of
this effect, as the J2 approximation used in this work
precludes an accurate calculation of the true frequency
of resonance hopping.
Another important avenue for future work exists in our
prediction that the eight TNOs considered in this work
populate two distinct dynamical classes: first, a class of
objects such as 2012 VP113 or Sedna, which are more
dynamically stable in the presence of large a, low e real-
izations of Planet Nine (which reduce closely to the case
of having no ninth planet); second, a class of objects that
are dynamically unstable in the presence of only Neptune
and no Planet Nine (and also in the case of a high-a, low-
e Planet Nine). This second class of objects may have
higher e and a than the first, and require a stabilizing
influence in the form of an eccentric Planet Nine to pre-
vent destabilizing interactions with Neptune. To truly
understand if this is a valid division, it is hoped that
a large number of high-a, trans-Neptunian-q, apsidally
aligned TNOs will be discovered in the next few years.
The classification of these objects, and exploration of the
mechanism by which Planet Nine may stabilize their or-
bits, should be explored in the future.
We thank Konstantin Batygin for a thorough review
of the manuscript and many useful suggestions. We
thank Andrew Vanderburg, Sarah Millholland, Jaehan
Bae, Marina Kounkel, and Ellen Price for useful con-
versations. We thank Michael Dieterle and Clara Eng
for helpful suggestions on how to effectively visualize the
simulation results. We also thank Gary Bernstein for re-
viewing the manuscript and suggesting several improve-
ments. Finally, we thank the referee for many useful
comments that improved the paper. This work was sup-
ported by NSF Grant AST-1515015. J.C.B and S.J.H.
are also supported by the NSF Graduate Research Fel-
lowship Grant No. DGE 1256260. The computations
for this work used the Extreme Science and Engineering
Discovery Environment (XSEDE), which is supported by
National Science Foundation grant number ACI-1053575.
This research was done using resources provided by the
Open Science Grid, which is supported by the National
Science Foundation and the U.S. Department of Energy's
Office of Science.
16
Becker et al.
REFERENCES
Alexandersen, M., Gladman, B., Kavelaars, J. J., et al. 2014,
Holman, M. J., & Payne, M. J. 2016, ArXiv e-prints,
ArXiv e-prints, arXiv:1411.7953
arXiv:1604.03180
Bailey, E., Batygin, K., & Brown, M. E. 2016, ArXiv e-prints,
Horner, J., Lykawka, P. S., Bannister, M. T., & Francis, P. 2012,
arXiv:1607.03963
MNRAS, 422, 2145
Bannister, M. T., Kavelaars, J. J., Petit, J.-M., et al. 2016, AJ,
Jones, E., Oliphant, T., Peterson, P., et al. 2001, SciPy: Open
152, 70
Batygin, K., & Brown, M. E. 2016a, AJ, 151, 22
-- . 2016b, ArXiv e-prints, arXiv:1610.04992
Becker, A. C., Arraki, K., Kaib, N. A., et al. 2008, ApJ, 682, L53
Beust, H. 2016, A&A, 590, L2
Brown, M. E., & Batygin, K. 2016, ApJ, 824, L23
Brown, M. E., Trujillo, C., & Rabinowitz, D. 2004, ApJ, 617, 645
Chambers, J. E. 1999, MNRAS, 304, 793
Chen, Y.-T., Kavelaars, J. J., Gwyn, S., et al. 2013, ApJ, 775, L8
Chen, Y.-T., Lin, H. W., Holman, M. J., et al. 2016, ApJ, 827,
L24
source scientific tools for Python, [Online; accessed 2016-10-24]
Lai, D. 2016, AJ, 152, 215
Lawler, S. M., Shankman, C., Kaib, N., et al. 2016, ArXiv
e-prints, arXiv:1605.06575
Li, G., & Adams, F. C. 2015, MNRAS, 448, 344
-- . 2016, ApJ, 823, L3
Lin, H. W., Chen, Y.-T., Holman, M. J., et al. 2016, AJ, 152, 147
Lykawka, P. S., & Mukai, T. 2007, Icarus, 189, 213
Malhotra, R., Volk, K., & Wang, X. 2016, ApJ, 824, L22
Mardling, R. A. 2013, MNRAS, 435, 2187
McKinney, W. 2010, in Proceedings of the 9th Python in Science
Dark Energy Survey Collaboration, Abbott, T., Abdalla, F. B.,
Conference, ed. S. van der Walt & J. Millman, 51 -- 56
et al. 2016, MNRAS, 460, 1270
de la Fuente Marcos, C., de la Fuente Marcos, R., & Aarseth,
S. J. 2016, MNRAS, 460, L123
Fienga, A., Laskar, J., Manche, H., & Gastineau, M. 2016, A&A,
Millholland, S., & Laughlin, G. 2017, AJ, 153, 91
Morbidelli, A., & Levison, H. F. 2004, AJ, 128, 2564
Nesvorn´y, D., & Roig, F. 2001, Icarus, 150, 104
Petit, J.-M., Kavelaars, J. J., Gladman, B. J., et al. 2011, AJ,
587, L8
142, 131
Gerdes, D. W., Jennings, R. J., Bernstein, G. M., et al. 2016, AJ,
Sheppard, S. S., & Trujillo, C. 2016, ArXiv e-prints,
151, 39
arXiv:1608.08772
Gerdes, D. W., Sako, M., Hamilton, S., et al. 2017, ApJ, 839, L15
Gomes, R., Deienno, R., & Morbidelli, A. 2017, AJ, 153, 27
Guan, P., Zhou, L.-Y., & Li, J. 2012, Research in Astronomy and
Towns, J., Cockerill, T., Dahan, M., et al. 2014, Computing in
Science and Engineering, 16, 62
Trujillo, C. A., & Sheppard, S. S. 2014, Nature, 507, 471
Astrophysics, 12, 1549
A. APPENDIX: PRECESSION EQUATIONS OF MOTION FOR TNOS
To consider the secular motion of the TNOs in the presence of Planet Nine, we can treat the problem with a coplanar
approximation, as done in Batygin & Brown (2016a). It is important to note that this is a very rough approximation,
as the TNOs and Planet Nine are expected to be inclined relative to the inner solar system and each other.
The Hamiltonian for this system, as used in Batygin & Brown (2016a), is:
H = − 1
4
GM
a
(1 − e2)−3/2
(cid:19)
(cid:18) mia2
i
M a2
4(cid:88)
i=1
(cid:20) 1
(cid:18) a
(cid:19)2 1 + 3e2/2
− Gm9
a9
4
a9
(1 − e2
9)3/2
− 15
16
(cid:18) a
(cid:19)3
a9
ee9
1 + 3e2/4
(1 − e2
9)5/2
cos (ω9 − ω)
(11)
when M is the mass of the central body, the subscript 9 denotes the properties of Planet Nine, the subscript i within
the summation in the first term denotes the four giant planets, and a lack of subscript denotes the TNO for which the
Hamiltonian is being written.
1 − e2, we can find the equation of motion for the argument of perihelion ω by taking
Using canonical variable =
the derivative of Equation 11, such that dω/dt ∝ dH/d. The result of this yields
√
(cid:21)
(cid:21)
∝
dω
dt
mia2
i (e2−1)−2a−2− 15
16
m9
e9
e
(1−9e2/4) cos(ω−ω9)− 3
4
m9
(cid:19)4 √
(cid:18) a
a9
1 − e2
9)5/2
(1 − e2
(cid:19)3 √
(cid:18) a
a9
1 − e2
9)3/2
(1 − e2
, (12)
(cid:20) 3
4(cid:88)
4
i=1
when leading constants have been dropped, since in this work we care only about the relative contributions of the
terms of the equation of motion. The first term of Equation 12 represents the apsidal precession of a TNO with orbital
elements (a, e, ω) that is caused by the inner solar system (where the outer four giant planets are treated as a solar
oblateness and the terrestrial planets ignored). The latter two terms of Equation 12 include dependences on the mass
of Planet Nine (m9), eccentricity of Planet Nine (e9), semi-major axis of Planet Nine (a9), and argument of perihelion
of Planet Nine (e9). These two terms represent the apsidal precession of the TNO due to Planet Nine's influence.
Alone, the influence of Planet Nine or of the inner solar system would lead to precession of each TNO's ω. Taken
together, the two precession terms can either boost precession rates or slow them. With the proper choice of orbital
elements for Planet Nine, dω/dt can be set to be zero, leading to no net precession relative to the Katti-Range vector
of Planet Nine's orbit. Such a situation could result in a selection of TNOs exhibiting orbits that remain in roughly
the same regime of parameter space over time, potentially leading to alignment like that observed in the TNOs in our
solar system.
For completeness, we note that Equation 12 can also be derived from the disturbing function using Lagrange's
planetary equations. We use the disturbing function as formulated in Mardling (2013), which is written to work in the
dimensions of energy. The disturbing function can be written as
R = − 1
4
GM m
a
(1− e2)−3/2
+
Gm9m
a9
1
4
(cid:19)
(cid:18) mia2
i
M a2
4(cid:88)
i=1
Planet Nine Stability Survey
(cid:34)
(cid:18) a
(cid:19)2 1 + 3/2e2
a9
(1 − e2
9)3/2
(cid:18) a
a9
− 15 e e9
16
(cid:19)3 1 + 3e2/4
(1 − e2
9)5/2
17
(cid:21)(cid:35)
,
cos(ω − ω9)
(13)
This equation has been simplified under the assumption m << mp << M . We can use Lagrange's planetary equations
to find the equation of motion analogous to that in Equation 12. Specifically:
dω
dt
=
mνa2e
dR
de
(cid:19)4 √
(cid:18) a
a9
1 − e2
9)5/2
(1 − e2
(14)
(cid:21)
,
(cid:19)3 √
(cid:18) a
a9
1 − e2
9)3/2
(1 − e2
(cid:20) 3
4(cid:88)
4
i=1
when ν is the orbital frequency of the TNO. Substituting Equation 13 into the relevant Lagrange equation (Equation
14) yields the full equation of motion:
dω
dt
=
m9ν
M
mia2
i (e2 − 1)−2a−2 − 15
16
m9
e9
e
(1− 9e2/4) cos(ω − ω9)− 3
4
m9
(15)
which is equivalent to the result presented in Equation 12.
It is very important to note that the preceding derivation makes several major approximations:
• We assume a coplanar system and neglect orbital inclination of all bodies,
• In the construction of the Hamiltonian and the disturbing function, we ignore all short-order, resonant terms4
For all of these reasons, this analytic result should be treated as approximate. To do the problem properly, it is
important to use numerical N-body simulations.
B. APPENDIX: EFFECTS OF ALLOWING ORBITAL ELEMENTS TO VARY IN N-BODY SIMULATIONS
In the set of N-body simulations that we describe in Section 2 and use to construct the posterior probability
distribution given in Section 3, we use a population of Planet Nine realizations with fixed inclination i = 30 degrees,
argument of perihelion ω = 150 degrees, and longitude of the ascending node Ω = 113 degrees. This was done
because the amount of uncertainty in each of these measurements would require a computationally unfeasible number
of integrations to well-sample the parameter space. Instead, we chose the approximate best values for each angle, as
reported and used in prior literature.
However, we can recreate the probability posterior presented in Figure 4 for an additional set of 1500 N-body
integrations, while allowing these orbital angles to vary, and examine the amount of difference between this new
posterior and the old one as a first test. To do this, we ran 1500 more numerical N-body integrations with the same
numerical properties as our other set (hybrid symplectic and Bulirsch-Stoer (B-S) integrator in Mercury6 (Chambers
1999), conserving energy to 1 part in 1010, replacing the three inner giant planets with a solar J2, including each TNO
with orbital elements drawn from observational constraints). In this set of integrations, we sampled a9 from a uniform
range between (400, 1200) AU, and e9 between (0,1). However, instead of fixing the orbital angles, we sampled from
normal distributions centered on the Batygin & Brown (2016a) estimates (i9 = 30 degrees, ω9 = 150 degrees, Ω9 =
113 degrees) with widths of 30 degrees in each case. Then, we integrated each of the 1500 realizations forward for 4.5
Gyr.
The results of this new, second suite of integrations is presented in Figure 10, which can be directly compared to
Figure 4 (which was constructed with our original set of simulations). Comparing the two figures presents three major
conclusions: [1] The main parameter space preferred in each set is similar, with high eccentricities being less preferred
and the range from 0.3-0.5 eccentricity, 600-1000 AU being good in both sets. [2] The overall survival probabilities are
lower for the case where Planet Nine's orbital angles are allowed to vary, which indicates that altering these values
too much decreases dynamical stability of the TNOs overall.
[3] The posterior presented in Figure 10, with varying
orbital angles, has less variation between peaks and valleys, which is the danger of adding additional layers of variation
(which we did by allowing three orbital angles to vary).
The reason that we do not use this set of simulations that allow initial orbital angles of Planet Nine to vary as
the main set in this paper is that the parameter space is not well-sampled. Additionally, the three additional free
parameters bring with them three additional sources of variations. The robust exploration of this parameter space is
outside the scope of this work.
4 and as shown in Section 5, this is probably dangerous to do
18
Becker et al.
Fig. 10. -- The overall stability posterior for the semi-major axis and eccentricity of Planet Nine, when Planet Nine's orbital angles were
allowed to vary over a normal distribution centered on the best-guess values. This posterior was constructed by taking a summation of the
posteriors for each individual object, including the six objects used in Batygin & Brown (2016a) and the two new high-a, low-e objects from
Sheppard & Trujillo (2016). As compared to Figure 4, which did not allow the orbital angles to vary, the survival probability is lower when
orbital angles are allowed to vary, suggesting that the best-guess values leading to more probable alignment also lead to more dynamically
stable configurations.
|
1802.05034 | 2 | 1802 | 2018-07-03T14:14:37 | Cometary impactors on the TRAPPIST-1 planets can destroy all planetary atmospheres and rebuild secondary atmospheres on planets f, g, h | [
"astro-ph.EP"
] | The TRAPPIST-1 system is unique in that it has a chain of seven terrestrial Earth-like planets located close to or in its habitable zone. In this paper, we study the effect of potential cometary impacts on the TRAPPIST-1 planets and how they would affect the primordial atmospheres of these planets. We consider both atmospheric mass loss and volatile delivery with a view to assessing whether any sort of life has a chance to develop. We ran N-body simulations to investigate the orbital evolution of potential impacting comets, to determine which planets are more likely to be impacted and the distributions of impact velocities. We consider three scenarios that could potentially throw comets into the inner region (i.e within 0.1au where the seven planets are located) from an (as yet undetected) outer belt similar to the Kuiper belt or an Oort cloud: Planet scattering, the Kozai-Lidov mechanism and Galactic tides. For the different scenarios, we quantify, for each planet, how much atmospheric mass is lost and what mass of volatiles can be delivered over the age of the system depending on the mass scattered out of the outer belt. We find that the resulting high velocity impacts can easily destroy the primordial atmospheres of all seven planets, even if the mass scattered from the outer belt is as low as that of the Kuiper belt. However, we find that the atmospheres of the outermost planets f, g and h can also easily be replenished with cometary volatiles (e.g. $\sim$ an Earth ocean mass of water could be delivered). These scenarios would thus imply that the atmospheres of these outermost planets could be more massive than those of the innermost planets, and have volatiles-enriched composition. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 1–28 (2002)
Printed 4 July 2018
(MN LATEX style file v2.2)
Cometary impactors on the TRAPPIST-1 planets can
destroy all planetary atmospheres and rebuild secondary
atmospheres on planets f, g, h
Quentin Kral,1(cid:63) Mark C. Wyatt,1 Amaury H.M.J. Triaud,1,2 Sebastian Marino, 1
Philippe Th´ebault, 3 Oliver Shorttle, 1,4
1Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
2School of Physics & Astronomy, University of Birmingham, Edgbaston, Birmingham B15 2TT, UK
3LESIA-Observatoire de Paris, UPMC Univ. Paris 06, Univ. Paris-Diderot, France
4Department of Earth Sciences, University of Cambridge, Downing Street, Cambridge CB2 3EQ, UK
Accepted 1928 December 15. Received 1928 December 14; in original form 1928 October 11
ABSTRACT
The TRAPPIST-1 system is unique in that it has a chain of seven terrestrial Earth-like
planets located close to or in its habitable zone. In this paper, we study the effect of
potential cometary impacts on the TRAPPIST-1 planets and how they would affect
the primordial atmospheres of these planets. We consider both atmospheric mass loss
and volatile delivery with a view to assessing whether any sort of life has a chance to
develop. We ran N-body simulations to investigate the orbital evolution of potential
impacting comets, to determine which planets are more likely to be impacted and the
distributions of impact velocities. We consider three scenarios that could potentially
throw comets into the inner region (i.e within 0.1au where the seven planets are
located) from an (as yet undetected) outer belt similar to the Kuiper belt or an
Oort cloud: Planet scattering, the Kozai-Lidov mechanism and Galactic tides. For the
different scenarios, we quantify, for each planet, how much atmospheric mass is lost
and what mass of volatiles can be delivered over the age of the system depending
on the mass scattered out of the outer belt. We find that the resulting high velocity
impacts can easily destroy the primordial atmospheres of all seven planets, even if
the mass scattered from the outer belt is as low as that of the Kuiper belt. However,
we find that the atmospheres of the outermost planets f, g and h can also easily be
replenished with cometary volatiles (e.g. ∼ an Earth ocean mass of water could be
delivered). These scenarios would thus imply that the atmospheres of these outermost
planets could be more massive than those of the innermost planets, and have volatiles-
enriched composition.
Key words: accretion, accretion discs atmospheres low mass stars (TRAPPIST-1)
circumstellar matter Planetary Systems.
8
1
0
2
l
u
J
3
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
4
3
0
5
0
.
2
0
8
1
:
v
i
X
r
a
1 INTRODUCTION
(d=12pc) M8V ultra-cool dwarf
The nearby
star
TRAPPIST-1 (2MASS J23062928-0502285) is now known
to be surrounded by at least seven terrestrial-like planets
(Gillon et al. 2016, 2017; Luger et al. 2017). This old
(7.6±2.2 Gyr, Burgasser & Mamajek 2017), close-by,
multi-planetary system may offer one of our best chances
to study the chemistry, and structure of terrestrial planet
atmospheres outside our Solar System (de Wit et al. 2016;
Morley et al. 2017). Moreover, several of the TRAPPIST-1
(cid:63) E-mail: [email protected]
c(cid:13) 2002 RAS
f and g, Gillon et al.
planets (most likely planets e,
2017) lie within the liquid water habitable zone (HZ,
e.g. O'Malley-James & Kaltenegger 2017). However, the
presence of liquid water and possible life strongly depends
on the atmospheric content of these planets, the presence
of oceans, the vegetation coverage, etc. (e.g. Wolf 2017;
Alberti et al. 2017; Ehlmann et al. 2016; Carone et al. 2016;
Godolt et al. 2016).
This system being very close-by, we may soon be able
to start characterising the atmospheres of the seven plan-
ets with new telescopes such as JWST (Barstow & Irwin
2016; O'Malley-James & Kaltenegger 2017) and the E-ELT
(Rodler & L´opez-Morales 2014; Turbet et al. 2016) and
2 Q. Kral et al.
search for tracers of life. Such detailed spectral characterisa-
tion may eventually allow us to infer the presence of biologi-
cal activity via the detection of gases such as ozone (Barstow
& Irwin 2016), or via the spectral signatures of pigmented
micro-organisms (Poch et al. 2017). Regardless, such ob-
servations will inform on the atmospheric compositions of
these planets that is necessary to study the possibility that
life may develop.
For now,
little is known about the atmospheres of
these seven planets. The two innermost planets b and c
have been observed using transmission spectroscopy (de Wit
et al. 2016). This showed that the combined spectrum of
both planets (obtained when transiting at the same time)
is featureless, which favours atmospheres that are tenous
(composed of a variety of chemical species), not hydrogen-
dominated, dominated by aerosols or non-existent. Similar
conclusions have been made for planets d, e, f (and poten-
tially g) by de Wit et al. (2018). Also, from the combined
measurement of planet radii (transit) and masses (transit
timing variations), the derived planets' densities show that
TRAPPIST-1 b, d, f, g, and h may require envelopes of
volatiles in the form of thick atmospheres, oceans, or ice
(Grimm et al. 2018). We thus do not know yet information
that would be important for considering the habitability of
the planets such as whether these planets' atmospheres are
primordial or created later, for instance by cometary im-
pacts, although current observations suggest that current
atmospheres may not be primordial due to a lack of hydro-
gen signatures in the observed spectra (de Wit et al. 2018).
Previous theoretical studies of the atmospheric compo-
sition of the TRAPPIST-1 planets have shown that they
may vary with time and be affected by the early evolution
of the star. Indeed, ultra-cool dwarfs such as TRAPPIST-1
take up to 1Gyr to cool down (Baraffe et al. 2015) and reach
the main-sequence after the planets formed. This means that
planets that are today in the HZ would have undergone a
very hot pre-main-sequence era (with potentially a runaway
greenhouse phase) and may have lost all (or part) of their
initial water content (Bolmont et al. 2017). Moreover, Bour-
rier et al. (2017) find that the total XUV emission from the
star might be strong enough to entirely strip the primordial
atmospheres of the planets over a few Gyr. One could then
expect that a few of the TRAPPIST-1 planets are devoid of
atmospheres, or left with a gas layer too tenuous for life to
persist over long timescales (Roettenbacher & Kane 2017).
Here we consider another process that can strongly in-
fluence the atmospheres, both positively and negatively for
life: exocomets. Impacting exocomets can influence planetary
atmospheres in multiple ways: a) they can provide an energy
source that depletes primordial atmospheres. b) They might
also deliver volatiles that subsequently replenish a secondary
atmosphere (i.e., dry, depleted atmospheres from impacts
or XUV irradiation could be replenished by later impacts,
and surviving primordial atmospheres could see their ele-
mental abundances significantly transformed via exocomet
impacts). c) Impacting exocomets may also act as catalysts
for the development of life. Indeed, these impacts may ini-
tiate a cascade of chemical reactions, some of which can
produce the necessary precursors to nucleobases on these
planets (Saladino et al. 2012; Ferus et al. 2015; Patel et al.
2015; Sutherland 2017; Ranjan et al. 2017).
For now, there is no evidence of exocomets in the
TRAPPIST-1 system, however, this does not mean they are
not present and part of the motivation of this work is to de-
termine if evidence for such a population may be imprinted
on the planets' atmospheres.
Many stars have large outer reservoirs of planetesimals
that produce a detectable infrared excess due to collisional
production of dust (Wyatt 2008; Eiroa et al. 2013). Detec-
tions of CO gas in several systems are used to infer that
these planetesimals are icy with a composition that is simi-
lar to Solar System comets (e.g. Kral et al. 2016; Marino et
al. 2016; Matr`a et al. 2017a). These planetesimal belts are
harder to detect around low mass stars such as TRAPPIST-
1 due to their low luminosity but this does not mean they are
not present (Plavchan et al. 2009; Theissen & West 2014).
Some stars also have evidence that comets from these outer
regions are being scattered into the inner regions. For exam-
ple, CO detected at 20au in η Corvi is inferred to originate in
the sublimation of such an exocomet population(Marino et
al. 2017). In addition, high-velocity metallic gas absorption
lines in some systems (Montgomery & Welsh 2012; Kiefer et
al. 2014; Eiroa et al. 2016) are inferred to originate in very
eccentric comets passing very close to their host star (called
falling evaporating bodies, e.g. Beust et al. 1990). Thus, it
is not unreasonable that TRAPPIST-1 has (or indeed may
have had) comets at some level.
In this study, we hypothesize that such comets exist
in the TRAPPIST-1 system and use previous studies that
looked at the effect of impacts onto planetary atmospheres
(e.g. de Niem et al. 2012; Schlichting et al. 2015), and espe-
cially hydrodynamical simulations (Shuvalov 2009) to derive
some constraints on the TRAPPIST-1 planets' atmospheres
in the presence of impacting comets.
We start by estimating the possible mass of a planetes-
imal belt that could have survived around TRAPPIST-1.
In Sec. 3, we then study the dynamics of comets in the
TRAPPIST-1 system that come close to the seven plan-
ets, i.e within 0.1au. Notably, we look into which planet
will receive most impacts, at which velocity and derive the
timescales on which impacts happen. In Sec. 4, we describe
three plausible scenarios that can potentially scatter many
exocomets over the lifetime of the system. In Sec. 5, we show
the results of our model, i.e how much atmospheric mass is
removed from the primordial atmospheres of the seven plan-
ets by a continuous series of impacts and evaluate whether
those impacts increase or reduce the amount of volatiles in
the planets' atmospheres, and what kind of atmosphere each
planet is likely to end up with. We then discuss our results
in terms of their implications for the development of life in
Sec. 6 before concluding.
2 THE POSSIBLE PRESENCE OF A DISC
AROUND TRAPPIST-1
This paper is based on the potential presence of a yet unde-
tected debris disc around TRAPPIST-1. To consider what
this debris disc might look like, we construct a minimum
mass extrasolar nebula for the TRAPPIST-1 system sim-
ilar to Hayashi (1981), or Chiang & Laughlin (2013) who
used 1925 extrasolar planets to constrain the minimum sur-
face densities at different distances from the star assuming
planets formed in situ.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Effects of impacts on the TRAPPIST-1 planets
3
yr (Wyatt 2008). This gives tc ∼ 4 × 103(r/1au)4.3yr by as-
suming typical values (as in Wyatt 2008; Kral et al. 2017c,
i.e e = 0.05, Q(cid:63)
D = 500J/kg, Dc=100km). In other words, a
belt at 1au would be significantly depleted after 7Gyr (the
age of the system) of collisional evolution and we expect any
belt this close in to have been significantly depleted. How-
ever, a belt at >28au could survive over 7Gyr. At shorter
radii, the mass that remains after collisional evolution for
7Gyr would be expected to have a radial profile that scales
∝ r7/3 (Kennedy & Wyatt 2010) as shown by the red dotted
line in Fig. 1. While this formula depends on many uncertain
parameters, it shows that we expect any potential surviving
belt to be located at (cid:38) 10au.
Using the extrapolation in Eq. 1, we expect such a left-
over belt between 10 and 50au to have a mass of ∼20M⊕,
which is compatible with the predicted large initial mass of
the protoplanetary disc around TRAPPIST-1 required to
have formed the seven planets (Haworth et al. 2018). While
this is at least two orders of magnitude more massive than
the Kuiper belt (Fraser & Kavelaars 2009; Vitense et al.
2010), note that the Kuiper belt is thought to have formed
much more massive, with a solid mass of 20-40M⊕ com-
patible with the MMSN (e.g. Weidenschilling 1977; Hayashi
1981; Nesvorn´y & Morbidelli 2012, but see Shannon et al.
2016 for a dissenting view). The left-over belt is not ex-
pected to extend much farther than 50-100au because pro-
toplanetary discs around low-mass stars are less extended
than around T-Tauri stars (Hendler et al. 2017). One caveat
to this estimate is that our approach is only accurate for
an in situ formation of the seven planets. For planets that
formed further out close to the water iceline as suggested by
Ormel et al. (2017), the surface density would go down by a
factor 10 at most and so would the belt mass leading to an
estimate of (cid:38)2M⊕.
The only observation of TRAPPIST-1 in the infrared is
by WISE at 22µm (Patel et al. 2017), which shows no signs
of infrared excess. However, any belt that is warm enough
to emit at 22µm would have to be inside 10au and so, as
noted above, would be expected to be collisionally depleted.
The only region where significant mass is expected to re-
main at 7Gyr is beyond 10au, where such a belt would be
< 15K (assuming a black body) and therefore its emission
would peak at λ > 340µm. This WISE observation is thus
not constraining and observations at longer wavelengths are
required to constrain such a cold belt, for instance using the
ALMA interferometer.
3 DYNAMICS OF IMPACTS FOR COMETS
COMING FROM AN OUTER BELT
There are many possible origins for the comets that may
impact planets b to h. Rather than studying the details of
the specific evolution for each scenario, we will start by as-
suming that very eccentric comets are produced and we will
study their dynamics and look at their interactions with the
seven planets. This framework is therefore general as soon
from the collisional cascade have a large enough collision velocity
that they can fragment after an impact. Depending on the level
of stirring, using this formula for radii (cid:38)50au is therefore not
accurate and only gives a lower limit on the timescale tc.
Figure 1. Surface density in the TRAPPIST-1 system assuming
a minimum mass extrasolar nebula and extrapolating to tens of
au to obtain a plausible mass that would be left in a potential,
yet undetected, belt. In red, we show the predicted profile after
7Gyr of collisional evolution.
To get a surface density for each planet, we take the
planet mass and divide it by the area of the annulus around
the planet. For planets c to g, we define the annulus as being
between the two midpoints to the neighbouring planets. For
planets b and h, we work out the half width using the centres
between planets b and c and between planets g and h and
multiply that width by two. This gives the following surface
density (in solids) after fitting the data (see Fig. 1)
(cid:16) r
(cid:17)−1.97
1au
Σ ∼ 122
kg/m2,
(1)
where r is the distance to the star. Our fit of Σ provides
values a factor 4 smaller than Chiang & Laughlin (2013)
at 1au (who used a large sample of Kepler planets around
earlier-type stars) but steeper in r and very close to the fit
by Gaidos (2017) who did it specifically for M-dwarf Kepler
planets. It is less than a factor 2 from the minimum mass
solar nebula (MMSN) in solids for terrestrial planets at 1au
(Hayashi 1981).
The H2O iceline during planetesimal formation is es-
timated to have been close to ∼0.1au in the TRAPPIST-1
system (Ormel et al. 2017). It could as well have been slightly
closer-in (by a factor ∼2) based on the (still not-well con-
strained) gradient of water compositions of the 7 planets
(Unterborn et al. 2018). We assumed that beyond 0.1au,
the solid {rock+ice} surface density is a factor 4 higher fol-
lowing Hayashi (1981). We can now extrapolate the mass
that may be present at several au and potentially form a
disc of planetesimals rather than planets.
A planetesimal belt at a radius r with dr/r ∼ 0.5 would
have a mass of ∼ 12.6(r/1au)0.03M⊕. The collisional life-
time of the biggest planetesimals in such a belt is given
by1 tc = 1.4 × 10−3 r13/3(dr/r)DcQ(cid:63)5/6
/Mtot
D e−5/3M
−4/3
(cid:63)
1 We note that this formula can be used when the largest bodies
c(cid:13) 2002 RAS, MNRAS 000, 1–28
10−210−1100101102Distancetothestar(au)10−210−1100101102103104105106107Surfacedensity(kg/m2)dataFitExtrapolationAfter7Gyr4 Q. Kral et al.
as eccentric comets are produced and will be tied to spe-
cific scenarios (planet scattering, Kozai-Lidov mechanism or
Galactic tides) in Sec. 4.
The pericentre q of the eccentric comets we model can
reach a few tenths to hundredths of au where they can collide
with one of the seven detected planets around TRAPPIST-
1. The apocentre Q can vary from a few au (for comets that
originate in close-in belts) to > 100s of au for comets coming
from very cold outer belts or exo-Oort clouds. We perform
N-body simulations of these very eccentric orbits assuming
that the evolution is dominated by perturbations from the
known TRAPPIST-1 planets to understand how their fate
depends on the comet's orbital parameters q (pericentre)
and Q (apocentre). That is, for each of these different comet
families (i.e for a given set of {q,Q}) we determine the frac-
tion that is accreted onto the different planets and the frac-
tion that is ejected. We also compute impact velocities for
each family of comets, which are used in Sec. 5 to assess if
cometary impacts are able to destroy planetary atmospheres
and if delivery of volatiles from these comets is possible.
3.1 N-body simulations of impacts with the seven
planets
The N-body simulations are run with REBOUND (Rein &
Liu 2012) with the Hermes integrator which combines the
IAS15 integrator for close encounters within a few Hill radii
of the planets (Rein & Spiegel 2015) and uses the WHFast
integrator otherwise (Rein & Tamayo 2015). The simulations
include the seven planets orbiting around the central star
TRAPPIST-1 (see Tab. 1 for the parameters used). We use
a timestep of 5% of planet b's orbital timescale. We assumed
zero eccentricities for the planets as the 2σ upper limits are
low (< 0.09 as implied by tidal forces and orbital stability,
Gillon et al. 2017; Tamayo et al. 2017; Quarles et al. 2017).
The planets gravitationally interact with each other, but
their orbits do not evolve significantly over the course of all
our simulations.
We start each simulation with 2000 test particles that
all have a similar pericentre and apocentre {q,Q} spread in a
narrow range defined by a grid (see Fig. 2). We run the sim-
ulations until all test particles have either been ejected from
the system (i.e., if their positions go beyond 100 times the
initial comet's apocentre) or accreted onto the planets or the
star. We note however that almost no particles collide with
the central star. This is because for high-eccentricity orbits
the pericentres will be locked and for low-eccentricity orbits,
we notice that there are very few scattering events that could
potentially send the comets onto the star. Rather, the parti-
cles tend to be accreted or ejected by the planet close to their
pericentres. We assumed zero inclination, which we expect
to be unrealistic but leads to much faster simulations and
can be scaled a posteriori to give results for a comet-like
inclination distribution (see subsection 3.5). Running the
simulations assuming a zero inclination angle was necessary
to allow the simulations to be performed in a reasonable
timescale (i.e., not exceeding two months). The whole set
of 900 simulations took ∼2 months on 20 CPUs, whereas
inclined comets would have taken almost two years to com-
pute. This is because we ran each simulation until there are
no particles left. As the time to accrete/eject particles is
much smaller in the zero inclination case, we gain a factor
greater than 10 in overall computational time. We note that
of the results we derive in this section, the probability map
as well as the accretion/ejection timescales are affected (in
a quantifiable way) by a change in inclination but not the
impact velocities (subsection 3.4).
We ran a grid of 900 N-body simulations for a wide
range of {q,Q} values, with 90 logarithmically-spaced bins
in pericentre covering 10−3au < q < 10−1au and 10
logarithmically-spaced bins in apocentre covering 10−1au <
Q < 102au, which form the grid seen in Fig. 2. The grid is
defined by the pericentres and apocentres at the start of the
simulations. The TRAPPIST-1 planets are located between
0.01 and 0.06au (white vertical lines in Fig. 2) so that the
chosen range of pericentres is large enough to follow what
happens when the comets' orbits cross those of the planets.
3.2 Probability to impact the different planets or
to be ejected for comet-like orbits
Fig. 2 shows a map of the probability to impact the differ-
ent planets (each inset is for a given planet, planet b to h
from left to right), while Fig. 3 shows the probability to be
ejected for each given {q,Q} of our parameter space. Some
of the large scale features in these figures can be readily
understood.
For example, the extended black regions in Fig. 2 at
large pericentres are because in order for a comet to collide
with a given planet, the comet's pericentre must be smaller
than the planet's semi-major axis apla. Since the pericen-
tre and apocentre of comets do not evolve significantly from
their starting values, this means that the region of the pa-
rameter space with q > apla appears in black. Comets with
such pericentres collide with the more distant planets.
Another large scale feature is that the probability to im-
pact one of the planets is higher for smaller cometary apoc-
entres. This can be explained by looking at Fig. 3 which
shows that the ejection rate goes up with increasing Q, not-
ing that the sum of the impact probabilities over the seven
planets and the ejection probability equals 1.
The increased ejection probability seen in Fig. 3 with
Q (for all pericentres) is because the comet's energy (∼
GM(cid:63)/Q) is lower for these larger apocentres and so a comet
is ejected by a smaller kick when passing by a planet. The
biggest kick in velocity that the comet can receive from a
planet (without colliding onto it) is roughly equal to vesc
(Wyatt et al. 2017), where vesc is the escape velocity of the
planet. The resulting increase in the comet's orbital energy
can be enough to unbind the comet if vorbvesc > GM(cid:63)/Q,
where the comet's orbital velocity vorb close to a planet is
(cid:115)
vcom ∼
2GM(cid:63)
apla
,
(2)
which leads to the apocentre value Q where ejections start
becoming dominant
(cid:18) M(cid:63)
(cid:19)1/2(cid:18) Mpla
0.08M(cid:12)
1M⊕
Q (cid:29) 0.2
(cid:19)−1/2(cid:18) Rpla
(cid:16) apla
1R⊕
(cid:19)1/2
(cid:17)1/2
0.03au
au,
(3)
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Table 1. Table describing the parameters used for the N-body simulations of the TRAPPIST-1 system (from Gillon et al. 2017; Luger
et al. 2017).
Effects of impacts on the TRAPPIST-1 planets
5
Star mass (M(cid:12))
Star radius (R(cid:12))
0.0802
0.117
Planets
Semi-major axis apla (10−3 au)
Radius Rpla (R⊕)
Mass Mpla (M⊕)
b
c
d
e
f
g
h
11.11
1.086
0.85
15.21
1.056
1.38
21.44
0.772
0.41
28.17
0.918
0.62
37.1
1.045
0.68
45.1
1.127
1.34
59.6
0.755
0.411
1 The mass of planet h is not well constrained (Luger et al. 2017) and as its radius is similar to
that of planet d, we assume the same mass as planet d.
Figure 2. Map showing the probability to impact (from left to right) planets b to h for each family of orbits defined by a given pericentre
q and apocentre Q. The vertical white lines show the positions of planets b to h. The white background colour is used to show the part
of the parameter space that have pericentres too far from the planets to either collide or be ejected. On the contrary, the black colour is
for orbits that collide with a low probability (cid:54) 10−2.
where Mpla, Rpla and apla are the planet mass, radius and
semi-major axis, respectively. This calculation explains why
Fig. 3 shows that for Q (cid:38) 1au, ejection is the more likely
outcome.
Another feature in Fig. 2 is that for pericentres inside
planet b, the accretion probability is higher for planets closer
to the star. In fact, the accretion probability decreases as
pla from planet b to h for a fixed {q, Q} in this regime.
−1
a
This can be explained by the different accretion timescales
of each planet as showed in Sec. 3.3.
Finally, another noticeable feature is that the high-
est probabilities of impacts (the narrow yellow regions) are
for comet orbits that have pericentres close to but slightly
smaller than the positions of the planets. For instance, on
the planet d inset, we see that the yellow region is concen-
trated in a narrow region of the parameter space between
0.015−0.021au (planets c and d positions). This can be read-
ily explained because comets with such a pericentre cannot
collide with planets b, c so increasing the rate of collisions
with planets d, e, f, g, h. The most extreme case is for comets
that have pericentres just below planet h, thus ensuring that
they can only collide with planet h and explaining the very
narrow yellow region in the planet h inset.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
3.3 Accretion/ejection timescales
It is important to consider the timescale on which particles
are accreted (or ejected) in the simulations, because we will
later be considering how these outcomes compete with other
processes that may be acting to modify the particles' orbits
(such as the processes that brought them onto comet-like
orbits in the first place).
In Fig. 4, we plot the loss timescale tloss which is the
timescale for half of the 2000 test particles to be lost from
the simulation (through accretion or ejection) as a function
of the apocentre Q. Since there is little dependence on the
pericentre of the comets' orbits (because the comet velocity
is almost independent of q), this shows that for Q (cid:28) 1au,
the loss timescale scales as Q3/2 and as Q1/2 for larger Q.
For Q (cid:28) 1au, the loss of particles is dominated by accre-
tion onto the planets. For a 2D geometry, the rate of colli-
sions between a given comet and planet is proportional to
Rcol = nσσ2Dvrel, where nσ is the fraction of the comet's
orbit per unit cross-section spent in a region dr around the
planet's orbit, vrel is the relative velocity at encounter and
σ2D is the collisional cross section. Considering the fraction
of the orbit spent in an annulus dr around the planet's orbit
we find that nσ (per cross-sectional area) is ∝ Q−3/2a
−1/2
.
pla
In practice, the velocity at encounter vrel is the same as
10−310−210−110−1100101102ApocentreQ(au)b10−310−210−110−1100101102c10−310−210−110−1100101102d10−310−210−1Pericentreq(au)10−1100101102e10−310−210−110−1100101102f10−310−210−110−1100101102g10−310−210−110−1100101102h10−210−1100ProbabilitytoimpactPlanet6 Q. Kral et al.
Figure 3. Map showing the probability to be ejected for an orbit
with a given pericentre and apocentre {q,Q}. The vertical green
lines show the positions of planets b to h. The white colour is used
to show the part of the parameter space that have pericentres
too far from the planets to either collide or be ejected. The black
colour is for orbits that collide and eject particles but with a very
low ejection probability (cid:54) 10−2.
Figure 4. Plot showing the removal timescale (in yr) for a comet
as a function of the apocentre Q (in au) for an i = 0◦ (crosses)
and a realistic comet-like inclination distribution (filled dots, see
subsection 3.5).
Figure 5. Accretion timescale tacc = R−1
col as a function of the
semi-major axes of the planets apla for an i = 0◦ (crosses) and a
realistic comet-like inclination distribution (filled dots). We show
the analytically predicted tacc (blue line) for an inclined distri-
bution of comets from the i = 0◦ case (see subsection 3.5). Here,
tacc is plotted for planet b for Q ∼ 1au and it scales as Q3/2.
the impact velocity vimp, and we show in Sec. 3.4 that vimp
is close to the comet's velocity (see Eq. 2), which is large
enough for gravitational focusing to be ignored such that
σ2D = 2Rpla. Therefore, we find that Rcol ∝ Q−3/2a
−1
pla, so
that the accretion timescale is tacc ∝ R
col ∝ Q3/2apla, ex-
−1
plaining why the loss timescale scales as Q3/2 for small Q.
It also shows that the accretion timescale scales as apla as
shown in Fig. 5, where we plot tacc for planet i by computing
tloss/pi, where pi is the probability to be accreted on planet
i (that we have for every {q, Q} cell in Fig. 2). This also
explains why the accretion probability (∝ t−1
acc) decreases as
−1
a
pla from planet b to h as noted in Sec. 3.2.
For Q (cid:29) 1au, the loss is dominated by ejections. In
that case, the cross section σej used to calculate the rate of
ejection is proportional to the impact parameter bej at which
encounters are just strong enough to cause ejection. The
kick ∆v that the comet receives from a planet after a close
encounter scales with 1/b, and for the ejection to happen
vcom∆v > GM(cid:63)/Q (see Sec. 3.2). This means that for a
flat geometry σej ∝ Q and so tej ∝ (nσσejvrel)−1 ∝ Q1/2,
explaining the dependencies.
3.4 Impact velocities for the different planets
An important parameter to determine the effects of a
cometary impact onto the atmosphere of a planet is the im-
pact velocity.
In Fig. 6, we show histograms of impact velocities for
the different planets. We computed the impact velocities for
each simulation (i.e. for specific pericentres and apocentres),
c(cid:13) 2002 RAS, MNRAS 000, 1–28
10−310−210−1Pericentreq(au)10−1100101102ApocentreQ(au)10−210−1100Probabilitytobeejected10−1100101102Q(au)100101102103104tloss(yr)∝Q3/2∝Q1/2i=0◦comet-likeinclination10−2apla(au)102103104105106107108tacc(yr)∝a2pla∝aplai=0◦comet-likeinclinationPredictionEffects of impacts on the TRAPPIST-1 planets
7
but find that the distributions of impact velocity do not de-
pend significantly on the comet's pericentre. To get Fig. 6,
we therefore average the vimp distributions over the pericen-
tres in the grid, assuming that comet orbits are uniform in
log q (keeping a fixed apocentre). Averaging in this way re-
sults in more accurate histograms of impact velocities. To do
so, the impact velocities from the different simulations are
weighted by the probability to impact the different planets
(using Fig. 2). Furthermore, Fig. 7 shows that the medians
of the impact velocity distributions for each planet also do
not depend significantly on apocentre. Thus while Fig. 6
shows the distributions for an apocentre of ∼1au, these dis-
tributions are also representative of that of a large range of
apocentres.
We see that the impact velocity distribution is peaked at
a different location for each planet from ∼ 15 to ∼110km/s
from planet h to b. A much smaller secondary peak can
also be seen for each planet. This is because there are two
extreme types of impacts. Collisions can occur when the
comet is on a near radial orbit approaching or receding
from pericentre. They may also occur when the planet and
comet velocities are parallel (i.e when the comet encoun-
ters the planet near its pericentre). As shown by Eq. 2, the
comet velocity at impact is ∼(cid:112)2GM(cid:63)/apla (for apla (cid:28) a),
locity of (cid:112)GM(cid:63)/apla (which varies from ∼35km/s for the
(cid:112)GM(cid:63)/apla(
maximal for radial encounters at (cid:112)3GM(cid:63)/apla. We note
which is thus always higher than the planet's Keplerian ve-
farthest to ∼80km/s for the closest planet). Therefore, we
find that the impact velocity distributions should peak at
2 − 1) (33, 29, 24, 21, 18, 16, 14km/s, for
planet b to h) for parallel orbits at impact and would be
√
that the impact velocities are much greater than the escape
velocities of the planets (∼10km/s) and therefore gravita-
tional focusing is not important.
Thus the high velocity peaks correspond to comets col-
liding on radial orbits and the low velocity peaks to comets
falling on the planets at their pericentres (i.e., parallel col-
lision). By looking at Fig. 2, we see that for planet h, the
highest impact probability region (the yellow region) is very
narrow and restricted to comets whose pericentres are close
to planet h's position so that most collisions are going to be
parallel. This explains why the low velocity peak is higher
for this planet. For planet b, however, the yellow region is
large and not peaked close to planet b's semi-major axis.
Therefore, most impacts will happen with comets on nearly
radial orbits and the high velocity peak is therefore higher
than the low velocity peak. Histograms for the other planets
can be understood following the same procedure.
The non-dependence of impact velocities on apocentres
shown in Fig. 7 also derives from the velocity at impact,
which, as shown by Eq. 2, only depends on apla and not a.
We notice that the median velocities are close to the Keple-
rian velocities of the corresponding planets (Fig. 7).
3.5 Simulations for realistic inclinations
The simulations assumed comets with zero inclination. To
check how our results change for different inclinations, we
ran a set of 30 additional simulations (spread across the
{q, Q} parameter space) with more realistic comet-like in-
clinations. The chosen inclination distribution follows a
Rayleigh distribution peaking at 10 degrees, i.e close to the
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Figure 6. Histogram of impact velocities (cut in 100 bins) for
each planet weighted by the impact probability from Fig. 2. The
y-axis shows the percentage of impacts per velocity bin.
distribution of JFC2 comets (Levison & Duncan 1997; Di
Sisto et al. 2009). We find that the loss timescale (see Fig. 4,
filled dots) and the timescale for accretion onto the differ-
ent planets are affected (see Fig. 5, filled dots), but that the
impact velocities are unaffected.
The difference in tacc between the inclined and flat cases
in Fig. 5 can be explained by generalising the analytics in
subsection 3.3. The ratio of the rates at which comets col-
lide with a planet is expected to be (nvσ3D)/(nσσ2D) =
πRpla/(4Imaxapla), where nv is now number per unit vol-
ume in the vicinity of the planet and Imax the median incli-
nation of the comets in the 3D case. We plot this analytical
prediction (blue line) together with some numerical simula-
tions for a distribution of inclinations (filled dots) in Fig. 5.
A similar comparison shows that the dependence in Q re-
mains the same for tacc for the two types of simulations. We
thus conclude that the zero inclination simulation collisional
rates can be scaled to account for the inclined case.
To recover the probability map shown in Fig. 2 for the
case of an inclined distribution, we also need to rescale the
loss timescale tloss shown in Fig. 4 because the probability
pi to be accreted on a given planet i is equal to tloss/tacc.
The results for tloss from the inclined numerical simulations
are shown in Fig. 4 (filled dots). We can predict the change
in ejection timescale (which dominates tloss at Q (cid:29) 0.2au)
for the 3D case from the 2D simulations in the same way
as for the timescales for accretion onto the planets. This
prediction is that ejection timescale should be longer by
0.8(M(cid:63)/Mpla)(aplaImax)/Q. This is reasonably accurate but
2 The Jupiter-family comets are short period comets that orbit
part of the time in inner regions of our Solar System and whose
orbits are primarily influenced by Jupiter's gravity.
020406080100120140Impactvelocity(km/s)0.00.51.01.52.02.53.0PercentageofimpactspervelocitybinHistogramofimpactvelocitiesPlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth8 Q. Kral et al.
Figure 7. Median velocity of the distributions shown in Fig. 6
for different apocentres Q. The thin horizontal lines show the
Keplerian velocity of each planet for comparison.
we prefer to use the numerical ratio of tloss for the inclined
and zero-inclination cases which is best fit by a power law
equal to 63 (Imax/10◦) Q−0.61.
Using these different scalings, we calculate the new
probability map (see Fig. 8) and use that the sum of the
probabilities to be accreted onto each of the planets and
to be ejected equals 1 to compute a new ejection map (see
Fig. 9). Comparing the predictions from our scalings to the
different results from the inclined distribution simulations,
we find that we are accurate within a factor 2.
4 DIFFERENT SCENARIOS TO MAKE
ECCENTRIC COMETS
We have studied the dynamics of highly eccentric comets in
the presence of the seven TRAPPIST-1 planets in the previ-
ous section. Here, we consider three different scenarios that
can send comets from the outer regions of the TRAPPIST-1
system onto such eccentric orbits (see Fig. 10); 1) A plan-
etesimal disc is perturbed by a nearby planet and comets are
scattered inwards by this single planet or through a chain
of planets (similar to comets scattered in our Solar System,
e.g. Duncan & Levison 1997; Bonsor et al. 2012; Marino et
al. 2018). 2) A distant companion to TRAPPIST-1 forces
comets in a Kuiper belt-like disc to undergo Kozai-Lidov
oscillations (e.g. Nesvold et al. 2016), which can bring the
comets to very close pericentres. 3) Galatic tides perturb
a far away exo-Oort-cloud and send comets to decreasing
pericentres.
We assume that the evolution of comets' orbits in these
three scenarios can be approximated as an evolution in
which their apocentres Q remain constant and their peri-
centres q decrease at a constant rate q. This approximation
allows us to use the results of Sec. 3 to consider the outcome
for comets scattered into the inner regions without having
to consider the detailed dynamics of the comets' origin. The
simplified dynamics allows us to study a wide range of dif-
ferent possible scenarios. Owing to this simplification, the
results are expected to give order of magnitude correct esti-
mates, which is justified by the uncertainties on the presence
of a belt in this system and its yet unknown properties. We
explore expectations for the different q values for each of
these three scenarios below.
4.1 Impacts from comets scattered by a single or
a series of planets
In our Solar System, comets from the Kuiper belt are thrown
into the inner Solar System thanks to a series of scattering
by different planets. This planet scattering scenario has been
invoked multiple times (Nesvorn´y et al. 2010; Booth et al.
2009; Bonsor et al. 2014) to try to explain the presence of
hot dust around many stars (see Kral et al. 2017a, for a
review). More recently, Marino et al. (2018) studied the ef-
fect of scattering by a chain of planets for a large parameter
space so as to understand which planetary systems are more
suited to create large hot dust levels or maximise impacts on
the chain of planets. They ran simulations for 1Gyr for dif-
ferent chains of planets with semi-major axes ranging from 1
to 50au and planet masses ranging from a few to 100M⊕. In
our case, we consider that interactions with the innermost
planet of the chain dominate the comets' dynamical evolu-
tion as they reach to the very small pericentres considered
here. However, we also have to consider that some fraction
of comets would have been ejected or accreted by the other
planets before reaching the innermost planet of the chain.
Marino et al. (2018) show that for a wide range of planet
chain architectures, fin = 1 − 7% of comets originating in
an outer belt end up reaching the inner system. For a close-
in belt similar to the debated (see MacGregor et al. 2018)
belt recently invoked around the M-dwarf Proxima Centauri
(Anglada et al. 2017), a single planet at 1au could be enough
to scatter comets into the inner regions (but we note that
we showed in Sec. 2 that such a close-in belt is not likely
to have survived around TRAPPIST-1 unless it formed re-
cently). In that case, no comets are lost on the way through
the chain and fin = 1. We only consider the case of a planet
coplanar to the 7 planets as this system seems well-aligned.
A non-coplanar configuration would lead to an increased in-
clination distribution, which effect could be quantified using
the analytics from Sec. 3.5.
Conservation of the Tisserand parameter3 (Murray &
Dermott 1999) means that comets being scattered by an
innermost planet that is on a circular orbit can only reach
down to a minimum pericentre (see Bonsor et al. 2012).
Thus to scatter comets to small enough pericentres to reach
the TRAPPIST-1 planets' locations, we consider that the
innermost planet must be on an eccentric orbit, since in
that case there is no minimum pericentre constraint (see also
Frewen & Hansen 2014). We also consider that this planet
3 This parameter is a constant in the circular three-body prob-
lem, which constrains the orbit of a comet after being scattered
by a planet.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
10−1100101102ApocentreQ(au)050100150200Impactvelocity(km/s)PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlanethEffects of impacts on the TRAPPIST-1 planets
9
Figure 8. New probability map: similar to Fig. 2 but for an inclined distribution of comets (see Sec. 3.5 for more details).
Figure 9. New ejection map: similar to Fig. 3 but for an inclined
distribution of comets (see Sec. 3.5 for more details).
should not be too massive, so as not to eject the comets
before they can reach the innermost parts of the system.
Guided by the results of Frewen & Hansen (2014) we
consider 1 and 10M⊕ planets with a 0.4 eccentricity orbiting
at 1au to be representative of the kind of planets that are
able to put comets on orbits that are capable of colliding
with the seven known TRAPPIST-1 planets. We note that
such planets are not massive enough and not close enough
to gravitationally disturb the orbits of the seven currently
known planets as can be checked directly from Read & Wy-
att (2016), so that the system of the seven inner planets stays
stable even in the presence of such an additional planet (see
also Quarles et al. 2017). We also note that these planet
masses agree with current mass upper limits by Boss et al.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Figure 10. Schematic of the different scenarios tested that could
potentially scatter comets in the inner regions of the TRAPPIST-
1 system.
(2017) (i.e < 4.6MJup within a 1 yr period, and < 1.6MJup
within a 5 yr period). While an eccentricity of 0.4 is above
the median eccentricity found for Earth to Super-Earth mass
planets, such eccentric planets are observed. We also note
that our scenario would still work for lower eccentricities as
described in Frewen & Hansen (2014) but the q value would
vary.
Moreover, we find that the outermost planet interact-
10−310−210−110−1100101102ApocentreQ(au)b10−310−210−110−1100101102c10−310−210−110−1100101102d10−310−210−1Pericentreq(au)10−1100101102e10−310−210−110−1100101102f10−310−210−110−1100101102g10−310−210−110−1100101102h10−310−210−1100ProbabilitytoimpactPlanet10−310−210−1Pericentreq(au)10−1100101102ApocentreQ(au)10−210−1100Probabilitytobeejected0.06au0.01au1au100auSingle planetPlanetscatteringTRAPPIST-1 planetsPlanet chainKozaimechanismOuter companion10000auGalactic tides10 Q. Kral et al.
ing with the belt and causing the scattering could migrate
outwards if it has a mass (cid:46) 10M⊕ (see Eq. 58 in Ormel et
al. 2012, where we used the surface density of the potential
surviving belt shown in Fig. 1) and stall if it is more mas-
sive. Such a migration is beneficial to sending more comets
inwards as shown in Bonsor et al. (2014) because more
time is available for the scattering process and it can access
more material to scatter from. However, for a planet mass
(cid:46)0.1M⊕, the scattering would not be efficient anymore as
shown by Eq. 52 of Ormel et al. (2012). Too massive an outer
planet would also prevent material from being scattered in-
wards as it would be more likely ejected but disc evolution
models find that having another Jupiter in the TRAPPIST-
1 system is unlikely (Haworth et al. 2018). Marino et al.
(2018) show that for planet masses (cid:46)100M⊕, (cid:38) 2% of the
scattered comets still reach the inner region. More massive
planets such as a Jupiter would more likely eject most of the
material (see Wyatt et al. 2017).
We ran N-body simulations to follow the evolution of
test particles initially randomly located in the chaotic zone
(where resonances overlap, as classically defined in Wisdom
1980) of such an eccentric planet. The planet is located at
1au and simulations are run until the planet runs out of
material to scatter among the initial 5000 particles in the
chaotic zone. Fig. 11 shows the evolution of the distribu-
tion of pericentres of particles that have their apocentres
at the 10M⊕ planet location, which decreases steadily with
time. Quantifying the rate of this decrease by looking at the
evolution of the median of the distribution, we find that
qP ∼ 5 × 10
−5au/yr,
(4)
over many orbits for the 10M⊕ case. Running another simu-
lation for an Earth mass planet with similar eccentricity at
the same location, we find that qP ∼ 10−5au/yr.
While the path of an individual comet could be some-
what stochastic through the parameter space (jumping in
individual scattering events), the effect for an ensemble of
comets is that of a slow inward migration of q. Therefore,
we model this population, and how it is depleted due to in-
teractions with the seven inner planets, by assuming that
comets have an apocentre Q that is fixed at the position of
the innermost planet of the chain (1au), and considering the
various depletion pathways as the comets cross the param-
eter space in Figs. 2 and 3 at a constant rate q. That rate
depends on the mass of the planet, so we keep this as a free
parameter, noting that Eq. 4 gives realistic values. That rate
has a strong influence on the outcome (see Sec. 5).
4.2 Impacts from comets undergoing Kozai-Lidov
oscillations due to an outer companion
The incidence of binaries around M-dwarfs is around 27%
(Janson et al. 2012). Comets located at tens of hundreds
of au (either in a disc or a more spherical Oort-cloud like
distribution) could be perturbed by such distant compan-
ion stars. If the mutual inclination i0 of some comets with
this companion is greater than 39.23 degrees, the so-called
Kozai-Lidov cycle can start and the mutual inclination starts
decreasing while the eccentricity of the comet increases to
reach a maximum (Kozai 1962; Lidov 1962).
Figure 11. Distribution of pericentres at 3 epochs (80, 1180
and 4980yr) due to perturbations of an eccentric (e=0.4), 10M⊕
planet located at 1au. Particles are initially placed in the planet's
chaotic zone and the ones with apocentres at the planet get pe-
riodic kicks that push their pericentres q inwards. The median of
distributions are shown as vertical lines. The median of q goes
inward with time and we find q = 5 × 10−5au/yr for this case.
(cid:112)1 − (5/3) cos2(i0) (Innanen et al. 1997). This means that
For the case of a circular outer companion, the max-
imum eccentricity reached by the comets is given by
to reach a pericentre q < q1 = 10−2au (to be able to reach
the seven planets), the initial mutual inclination should be
(cid:16)(cid:113) 3
(cid:17)
5 (1 − (1 − q1/a)2)
greater than i0 > arccos
, where a is
the semi-major axes of the comets. If the belt is really close-
in, i.e a = 0.1au, this corresponds to i0 > 70.3 degrees, and
at 100au, it gives i0 > 89.4 degrees (i.e., an almost perpen-
dicular orbit is necessary in this latter case). We note that
this inclination is between the perturber and comets and the
latter can be inclined compared to the planets. Therefore,
finely tuned companions would be needed to send comets to
the right location.
However, for an eccentric outer companion, the comets'
eccentricity can reach values arbitrarily close to 1 (e.g. Lith-
wick & Naoz 2011; Teyssandier et al. 2013). While the peri-
astron precession (due to GR) can dominate the dynamics
(Liu et al. 2015) and stop the Kozai mechanism from work-
ing when the eccentricity comes close to 1, this occurs only
for pericentres interior to the known TRAPPIST-1 planets
(see Eq. 50 and Fig. 6 of Liu et al. 2015).
The Kozai oscillations will occur even if the perturbing
companion is very distant and/or not very massive, only the
timescale to achieve the eccentricity change will be longer in
that case. Assuming the initial eccentricities e0 of comets in
a disc are small then the timescale TK to reach the maximum
eccentricity given an eccentric perturber is (Antognini 2015;
Naoz 2016)
c(cid:13) 2002 RAS, MNRAS 000, 1–28
10−310−210−1100Pericentre(au)01020304050607080Distributionofpericentres(%)Evolutionofpericentresfora10M⊕eccentricplanetat1auTime=80yrTime=1180yrTime=4980yrEffects of impacts on the TRAPPIST-1 planets
11
(cid:18)
1 +
M(cid:63)
Mc
(cid:19)(cid:18) P 2
c
P
(cid:19)(cid:0)1 − e2
c
(cid:1)3/2
,
TK ∼ 2.7√
oct
(5)
where ac, ec, Mc and Pc are, respectively, the semi-major
axis, eccentricity, mass and orbital period of the companion
and P is the orbital period of the comet being perturbed (i.e
P = 2π(cid:112)a3/(GM(cid:63))). The parameter oct = ec(a/ac)/(1 −
e2
c) quantifies the relative size of the octupole term of the
Hamiltonian compared to the quadrupole term and is not
equal to zero for an eccentric perturber (the timescale for a
circular orbit can be found by substituting oct ∼ 259, see
Antognini 2015).
Then, we can determine an order of magnitude for the
rate of pericentre evolution given by qK ∼ (a(1 − e0) −
q1)/TK ∼ a(1 − e0)/TK , so that
qK ∼ 2 × 10
(cid:16) ec
0.4
−4(cid:16) a
(cid:17)1/2(cid:18) 1 − e2
100au
1 − 0.42
1 − 0.1
(cid:17)3(cid:18) 1 − e0
(cid:19)(cid:16) ac
(cid:19)−2(cid:18) 1 + M(cid:63)/Mc
(cid:18) M(cid:63)
(cid:19)1/2
1 + 0.08/0.01
150au
c
(cid:17)−7/2
(cid:19)−1
0.08M(cid:12)
au/yr.
(6)
Therefore, considering a belt at 100au perturbed by an ec-
centric companion of mass 0.01M(cid:12) at 150au4, we find that
qK ∼ 2× 10−4au/yr. While a much farther companion could
decrease that value and a farther exo-Kuiper belt would in-
crease qK , we consider in Sec. 5.1 how evolution at typical
q might affect the planetary atmospheres of TRAPPIST-1
planets.
We have also checked that for such a configuration the
Kozai mechanism cannot be suppressed by the precession
induced by unknown planets in the system. Imagining the
worst case scenario of the presence of a Jupiter-mass planet
at 10au in this system, the Kozai dynamics remains dom-
inated by the outer companion if the belt is located fur-
ther than ∼30au (Petrovich & Munoz 2017), which is as-
sumed here. We also looked at the effect of the seven known
TRAPPIST-1 planets on the Kozai mechanism. The bodies
that can reach these planets must be very eccentric and to
take that into account properly, we model the effect of these
planets as being an effective J2 (quadrupole moment) and
check whether the precession rate due to J2, i.e. ωJ2 is able
to counteract the precession due to Kozai. Using Eq. 35 in
Fabrycky & Tremaine (2007), we find that the effective J2 of
the 7 TRAPPIST-1 planets starts contributing and reduce
the maximum Kozai eccentricity for
J2 > 105(cid:16) a
(cid:18) Mc
100au
0.01M(cid:12)
(cid:17)−3(cid:18) 1 − e2
(cid:19)−1(cid:16) a7
1 − 0.42
(cid:17)5(cid:16) ac
(cid:19)(cid:18) M(cid:63)
(cid:18)
(cid:112)
150au
(
0.08M(cid:12)
1 − e2 +
c
(cid:19)−3/2
(cid:17)−2
(cid:19)(cid:0)1 − e2(cid:1)2
0.06au
1
2
)2 − 1
4
,
(7)
where a7
is
the
semi-major axis of
the outermost
4 We note that such a low mass companion at large distances is
not yet ruled out (Howell et al. 2016; Boss et al. 2017).
c(cid:13) 2002 RAS, MNRAS 000, 1–28
TRAPPIST-1 planet, and e is the eccentricity of the comet,
which for a belt of semi-major axis a should be 1− 0.01/a to
be able to reach the innermost planet or 1− 0.06/a to reach
the outermost one. Using Batygin & Morbidelli (2017), we
find that the J2 due to the 7 TRAPPIST-1 planets would
be ∼ 2 × 10−5. Therefore, from Eq. 7, we estimate that for
a 0.01M(cid:12) at 150au, the belt of planetesimals should be at
(cid:38) 10au to be able to reach the outermost planet or at (cid:38) 70au
to reach the innermost one.
We acknowledge that the change in inclination while
undergoing Kozai oscillations is not taken into account in
our previous general simulations shown in Sec. 3. However,
depending on the exact inclination of the companion com-
pared to the belt, we can quantify using the equations given
in Sec. 3.5 how it will affect the probability to be accreted
onto the planets, which scales as I−1
max.
4.3 Impacts from Oort-cloud comets perturbed
by Galactic tides
TRAPPIST-1 may have an Oort cloud, either because
comets were captured from their neighbouring stars' belts
at the cluster stage (Levison et al. 2010), or because comets
were scattered out by its planetary system (Tremaine 1993).
In our Solar System, Duncan et al. (1987) propose that left-
over comets between Uranus and Neptune would be thrown
onto more extended orbits by the two planets until they
reach a semi-major axis of ∼ 5000au where Galactic tides
change their angular momentum, therefore moving their pe-
riastron from reach of the planets. While planets that are ef-
ficient at forming Oort clouds need to have the right ranges
of mass and semi-major axis, which does not include the
known TRAPPIST-1 planets (e.g. Wyatt et al. 2017), other
(as yet unseen) planets in the system could have scattered
material into such an Oort cloud. Moreover, an Oort-cloud
forming planet does not necessarily need to be at this exact
location now as it could have migrated.
The same mechanism,
i.e. Galactic tides, which in-
creased angular momentum of leftover comets pumped up
by Uranus and Neptune and thus detaching the comet or-
bits from the planets can also decrease angular momentum
and bring back an outer Oort-cloud comet to the planetary
system. For an Oort cloud, the Galactic tidal force (due to
the Galactic disc potential) will slowly make the comets lose
angular momentum resulting in a slow drift inwards of peri-
centre (because e increases, a is constant) at a rate qG (e.g.
Heisler & Tremaine 1986; Matese & Whitman 1992; Veras
& Evans 2013). The eccentricity reaches a maximum that
is given by Breiter & Ratajczak (2005), which is greater for
comets perpendicular to the orbital plane. It is usually as-
sumed that when a comet reaches a few au, it is lost from
the Oort cloud due to planetary perturbations (e.g. Heisler
& Tremaine 1986; Fouchard et al. 2006).
The value of qG can be estimated from the mean square
change in angular momentum per orbit (cid:104)∆J 2(cid:105) = 1.2 ×
10−29ρ2
0 a7/M(cid:63) (in au4/yr2, Eq. A4 of Wyatt et al. 2017),
with ρ0 the stellar mass density in units of 0.1M(cid:12)/pc3 (lo-
cal stellar mass density), a in au and M(cid:63) in M(cid:12). Thus, since
qG = (cid:112)(cid:104)∆J 2(cid:105) q/(2a3)/π, we estimate that (Duncan et al.
1987)
12 Q. Kral et al.
qG ∼ 7 × 10
−8
ρ0
(cid:18)
(cid:16)
(cid:19)(cid:16) a
(cid:17)1/2(cid:18) M(cid:63)
(cid:17)2
(cid:19)−1/2
0.1M(cid:12)/pc3
104au
q
0.06au
0.08M(cid:12)
au/yr.
(8)
We note that this qG value is very small but it varies strongly
with a.
Therefore, for the case of the TRAPPIST-1 planets, it
means that if the location of the Oort cloud is closer than a
few 103au, the time for moving the bulk of Oort-cloud bod-
ies down to small pericentres close to the planet positions
(i.e., < 0.06au) would be greater than the age of the system.
However, for an Oort-cloud at 105au, it only takes ∼5Myr to
reach the inner region but an origin at such a large distance
becomes unlikely given that such comets would have been
stripped by passing stars (Tremaine 1993). Given the age
and low-mass of TRAPPIST-1, comets with a semi-major
axis beyond 2000au should be strongly depleted by passing
stars but some may still remain. We also note that the pres-
ence of massive Jupiter-like planets in the outer regions of
the TRAPPIST-1 system would have strong effects on the
dynamics of the system (e.g. Kaib & Quinn 2009) and our
prescription would need to be revised if this type of planet
is discovered.
5 RESULTS
5.1 Fraction of comets that impact onto each
planet
5.1.1 For a general q
Here we use the results from sections 3.1 and 3.5 to deter-
mine how much material will be accreted on the different
planets depending on how fast comets move inwards, which
is assumed to be set by a constant rate of change of pericen-
tre q. For a given q and apocentre Q, each (pericentre) q cell
of our parameter space is progressively crossed as the comet
moves inwards to smaller q values. Taking into account the
timescales shown in Fig. 4 and scalings from Sec. 3.5 (as
we consider realistic inclined comets), we can use Figs. 2
and 3 (or their counterparts Figs. 8 and 9) to work out the
fraction of comets that are accreted onto the different plan-
ets or ejected along the way. Hereafter we only consider the
inclined case using the results from Sec. 3.5.
Fig. 12 shows the fraction accreted on the different plan-
ets facc for Q = 1 and 100au. For small q (i.e < 10−5au/yr
for the Q = 1au case) most comets end up on planets g
(yellow curve) and h (brown), while for large q each planet
gets a fraction of comets accreted. We also show the frac-
tion ejected as grey lines for Q = 1 and 100au, which is
close to 1 for very small
q and decreases for larger values,
meaning that comets can go past the planets without be-
ing ejected nor accreted onto the planets for q > 10−5au/yr.
These comets may end up on the star or collisionally deplete
before reaching it.
We see that the fraction of comets accreted5 facc onto
the different planets varies significantly from 0.05 to < 10−8
for 10−6 < q < 1au/yr. For small q, the fraction accreted is
dominated by planets h and g because q decreases so slowly
that these outermost planets catch all impacting comets be-
fore they reach further in. On the other hand, for large q,
the comets cannot efficiently accrete on the planets (as the
loss timescale is long compared to q/ q, see Fig. 4) and end
up at small radii (where they either accrete onto the star
or deplete collisionally). In between these two regimes, each
planet accretes a fraction of the scattered comets. The frac-
tion of comets accreted facc is also higher for smaller apoc-
entres, as expected from Fig. 2. We find that for large q,
facc ∝ q−1Q−1. The fraction accreted by the different plan-
ets vary by one order of magnitude in this regime (with b
and h representing the extremes). This is due to both the
difference in collisional cross sections and positions (since
tacc ∝ a2
pla, see Fig. 5). In subsection 5.1.2, we assess the
outcome for the specific values of q that have been derived
in Sec. 4 for the different scenarios.
5.1.2 For q derived from specific scenarios
In Fig. 12, we see that for a planet scattering scenario
(both for a single planet or a chain) in which Q = 1au
and q = 10−5au/yr (i.e. corresponding to a 1M⊕ planet in
Sec. 4.1), we end up in the regime where a fraction of comets
is accreted onto each planet. The fraction accreted is rather
high in this case (between 0.01 and 0.03 for planets b to
g) because the probability to be accreted for comets with
Q = 1au is rather high (as expected from Fig. 3). This frac-
tion accreted is valid for a single eccentric planet scattering
material from a close-in belt at ∼ 1au similar to the debated
belt potentially found around Proxima Cen (Anglada et al.
2017). However, comets coming from tens of au belts would
have to be scattered through a planet chain before making
it to the innermost planet of the chain and some will be lost
on the way. Here we assume that fin ∼ 5% of the comets
will make it to the innermost planet (see Sec. 4.1), which
reduces the fraction accreted on the different planets from
the initial reservoir (the Kuiper-belt like disc) to ∼ 5×10−4.
For the Kozai scenario, we consider that Q represents
the disc location from which the comets are perturbed by an
outer companion and we take a typical distance of 100au as
being representative. Considering the typical
qK value de-
rived (2 × 10−4 au/yr), Fig. 12 shows that we are in the
second regime where a fraction of comets (∼ 10−5) is ac-
creted onto each of the seven planets. This is close to two
orders of magnitude smaller than the chain of planets case.
Here, we do not have to reduce the number of comets that
arrives on to the seven planets (i.e fin = 1) as Kozai oscil-
lations operate directly from the outer belt.
For the Galactic tide scenario, from Fig. 12, we eval-
uate that an Oort cloud at a few 104au (that has a fast
enough qG to send comets to small pericentres within a
fraction of the age of the system), i.e with large apocen-
tres, the probability to be accreted is always (cid:46) 10−6 (for
all the plotted q, i.e more than order of magnitude smaller
than for the Kozai mechanism). Therefore, the fraction of
5 Here, we use accreted to say that a comet was not ejected but
instead impacted onto a planet. It does not mean that the entirety
of the accreted comet's material will stay in the atmosphere as
we will show later.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Effects of impacts on the TRAPPIST-1 planets
13
Figure 12. Fraction of comets accreted facc onto each planet as a function of q for Q = 1au (solid lines) and Q = 100au (dashed).
For large q, facc scales approximately as ∝ Q−1 so this plot can be used for all Q values. The Y-axis shows the fraction of comets that
hits the different planets (i.e., accretion regime). The grey lines show the ejected fraction of comets. The vertical black lines are typical
q values for two scenarios: planet scattering (with Q = 1au) and Kozai oscillations from a 100au belt (see the text in Sec. 5.1).
comets accreted is very low, which will not have any im-
pacts on the atmospheres. Therefore, we rule out Galactic
tides as being an efficient mechanism6 to modify the atmo-
spheres of the TRAPPIST-1 planets. We also note that the
same forces driving particles with high pericentres to low
pericentres could also drive back these low pericentre orbits
to high values, sometimes before they had time to reach the
7 inner planets (e.g. Emel'Yanenko et al. 2007; Rickman et
al. 2008), and thus makes this scenario even more unlikely.
We are, thus, left with two plausible mechanisms to throw
comets on the seven planets, namely, scattering by planets
and Kozai oscillations due to an outer companion.
6 We note that we could fine tune the position of the Oort cloud
to be in a narrow range in between 104 and 105au to maximise
the fraction accreted while allowing enough time for the comets
to reach the planetary system but this would always result in an
order of magnitude less efficient mechanism than Kozai. More-
over, it is not likely that an Oort cloud around a low mass star
such as TRAPPIST-1 forms farther out than in our Solar System
(Wyatt et al. 2017) as required here for maximising Galactic tide
effects. And as shown in Sec. 4.3, such distant belts should be
depleted owing to passing stars.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
5.2 The relative effect of different impactor sizes
on the atmospheres of the different planets
In the previous subsection, we analysed the fraction of
comets accreted facc by each planet. However, we want to
quantify the effect of these impacts on the atmospheres of
the seven planets. For example, we have seen in Fig. 7 that
impact velocities are much higher for planet d compared to
further planets, so even if the fraction accreted is the same
as that of the more distant planets in the planet scattering
scenario, the effect on atmospheric mass loss may still be
more important. Here, we quantify the atmospheric mass
loss, and projectile mass accreted in the atmosphere (rela-
tive to impactor masses), i.e., the impactor mass that does
not escape the atmosphere after impact, for the different
planets and for different impactor sizes.
We use the numerical study of the effect of impacts on
atmospheres by Shuvalov (2009) to derive some conclusions
for the TRAPPIST-1 planets. We first present the set of
equations from Shuvalov (2009) that we use to derive the
atmospheric mass loss and projectile mass accreted after a
given impact. The outcome depends on the dimensionless
variable η (Shuvalov 2009)
10−610−510−410−310−210−1100q(au/yr)10−810−710−610−510−410−310−210−1100AccretedfractiononPlanetsfaccQ=100auQ=1auPlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlanethEjection14 Q. Kral et al.
Figure 13. Left: Atmospheric mass loss to impactor mass ratio (Matmloss/Mimp)facc for a comet scattered from an outer belt for
the different planets as a function of impactor sizes D. It shows how much atmospheric mass is lost after a given comet is thrown in,
taking into account that the fraction of comets that hit the different planets is not equal to 1 as already seen in Fig. 12. Right: Accreted
projectile mass to impactor mass ratio (Mimpacc/Mimp)facc as a function of D. It shows how much projectile mass is accreted after a
given comet is thrown in. The different lines are for Q = 1 (solid lines) and 100au (dashed lines) with q = 10−5au/yr and for Q = 1au
and q = 10−1au/yr (dotted line).
(cid:18) D
(cid:19)3
ρprρt
imp − V 2
esc)
(V 2
H
ρatm0(ρpr + ρt)
V 2
esc
η =
,
(9)
where D is the impactor diameter, H the atmosphere scale
height (H = kT /(µmH g) for an isothermal atmosphere with
g = GMp/R2
p) and ρt, ρpr, ρatm0 are the densities of the
target (planet), projectile (exocomet), and atmosphere at
the surface, respectively. We assumed ρt = 5000 (terrestrial
planet-like), ρpr = 1200 (comet-like), ρatm0 = 1.2kg/m3,
µ = 28.97 (we assume an Earth-like atmosphere for now)
and T is taken to be the equilibrium temperature of the
planets (assuming a null Bond albedo as calculated in Gillon
et al. 2016). We note that recent observations suggest that
some of the TRAPPIST-1 planet densities may be slighlty
lower because of the potential presence of ice layers. Grimm
et al. (2018) find that water mass fractions < 5% can largely
explain the observed mass-radius relationship of the less
dense planets. Therefore, we can expect densities that are
10s of percent lower than assumed here, which would trans-
late as a small uncertainty on η, which is however much
lower than the uncertainties on the dynamics (see Sec. 4),
and is thus not considered here in details. Vimp is the impact
velocity and Vesc =(cid:112)2GMpla/Rpla is the escape velocity for
the different planets. We have seen in Sec. 3.4 that Vimp is
much greater than Vesc, which simplifies the previous and
following equations for most cases.
The atmospheric mass loss per impactor mass is then
defined as (Shuvalov 2009)
Matmloss
Mimp
=
imp − V 2
esc)
(V 2
V 2
esc
χa(η),
(10)
is
χa =
where Mimp
−6.375 + 5.239 log10 η − 2.121(log10 η)2 + 0.397(log10 η)3 −
0.037(log10 η)4 + 0.0013(log10 η)5.
the impactor mass and log10
To get meaningful results, we compare the atmospheric
mass loss Matmloss to the impactor mass (of size D) that
makes it to the inner regions and that is accreted on to the
planet. Therefore, using the previous notations, we are in-
terested in (Matmloss/Mimp)finfacc where we recall that fin
is the proportion of comets that are scattered from an outer
belt and make it to the inner regions and facc is the accreted
fraction onto a given planet. This ratio can therefore be un-
derstood as the atmospheric mass that is removed by one
comet scattered from an outer belt, where only "a fraction"
of the comet makes it to the inner regions and a fraction
of that is accreted onto a specific planet. In Fig. 13 (left),
we plot (Matmloss/Mimp)facc keeping in mind that we should
multiply that value by fin (if it is different than 1, see Ta-
ble 2) to get the real value of accreted comets that make it
to the inner regions.
Fig. 13 (left) shows (Matmloss/Mimp)facc as a func-
tion of impactor diameter D for Q = 1 and 100au (and
q ∼ 10−5au/yr) and for a higher q (10−1au/yr) and Q = 1au,
using the impact velocity distributions shown in Fig. 6. The
overall shape of the curves in Fig. 13 (left) is explained in
Shuvalov (2009). Impactors of size a few kms are the most
harmful at removing atmospheric mass. Impactors smaller
than 100m do not create large impact plumes and cannot
accelerate large atmospheric mass to high latitudes. For im-
pactors larger than a few 10s of kms, atmospheric erosion
continues to grow very slowly but the mass an impact re-
moves cannot be greater than the total local atmospheric
mass available. Therefore, for large impactors the atmo-
spheric mass removed per increasing impactor mass becomes
smaller.
The most harmful impactor size shifts along the x-axis
for the different planets mainly because of the change in im-
pactor velocity and the different properties of the planets
through H (the atmosphere scale height) with the relative
scalings given in Eq. 9. The variations along the y-axis are
mainly due to the different fraction accreted facc for each
different planet (see Fig. 12) and the different impact veloc-
ities (see Fig. 6) and scale as shown in Eq. 10. For example,
we see that even though planets d, e, f, g accrete at the same
level (see Fig. 12), the atmospheric mass loss is greater for
c(cid:13) 2002 RAS, MNRAS 000, 1–28
10−1100101102ImpactordiameterD(km)10−810−710−610−510−410−310−210−1100101102faccMatmloss/Mimp[q=10−5,Q=100][q=10−5,Q=1][q=10−1,Q=1][q=10−5,Q=100][q=10−5,Q=1][q=10−1,Q=1]PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth10−1100101102ImpactordiameterD(km)10−810−710−610−510−410−310−210−1100101102faccMimpacc/Mimp[q=10−5,Q=100][q=10−5,Q=1][q=10−1,Q=1][q=10−5,Q=100][q=10−5,Q=1][q=10−1,Q=1]PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlanethEffects of impacts on the TRAPPIST-1 planets
15
planet d because impact velocities are higher for the closer
in planets (see Fig. 6).
The effect of increasing Q from 1 to 100au (solid to
dashed lines) is to shift all the lines down by a factor 100
because facc decreases by a factor 100. Changing q from
10−5 (solid) to 10−1au/yr produces a shift downwards of
four orders of magnitude since facc decreases by a factor 104
between these two cases. Fig. 13 (left) can therefore be used
to work out the relative effectiveness of comets at removing
mass from the atmosphere of each planet for any given q and
Q, even though we show the results for only two different Q
(i.e., it is a general plot, not tied to a specific scenario from
Sec. 4, and only facc ∝ q−1Q−1 changes for different values
of q and Q, making it easy to compute results for different
q and Q).
The simulations of Shuvalov (2009) also showed that
the projectile mass accreted per impactor is given by
of ice by mass. For an asteroid-like body, the water mass
fraction is lower and is found to vary between 10−3 and 0.1
(Abe et al. 2000). We will assume an intermediate value
of 1% for asteroid-like bodies7, which is typical of ordinary
chondrites in our Solar System (but we note that carbona-
ceous chondrites can reach 10% of water by mass, Raymond
et al. 2004). This gives us two extreme volatile delivery sce-
narios to consider with our model.
The CO or H2O content of exocomets can be probed for
the most massive belts and are found to be similar to Solar
System comets (e.g. Kral et al. 2016; Marino et al. 2016;
Matr`a et al. 2017a). The potential to detect gas in debris
disc systems will improve with new missions (see Kral et
al. 2017b) and the assumptions used in this study could be
refined with future estimates of the volatile content of exo-
comets in the TRAPPIST-1 system to get a better handle
on the final atmospheric composition.
Mimpacc
= 1 − χpr(η),
Mimp
(11)
where χpr = min{1, 0.07(ρt/ρpr)(Vimp/Vesc)(log10 η − 1)}.
Similarly to atmospheric mass loss, Fig. 13 (right) shows
the accreted projectile mass per comet (Mimpacc/Mimp)facc
as a function of impactor diameter D for Q = 1 and 100au
(and q ∼ 10−5au/yr) and for a higher q (10−1au/yr) and
Q = 1au.
The shape of the curves in Fig. 13 (right) is already
known from Shuvalov (2009). The ejecta from impacting
bodies that are (cid:46) 1km does not have enough energy to es-
cape after impact and is stranded in the atmosphere (though
some material may condense on the planet surface at a later
point, see Sec. 6.3). For more massive bodies, the ejecta after
impact is increasingly more energetic until the airless limit is
reached (i.e., when atmospheric drag can be neglected before
the after-impact plume expansion) where all the projectile
material escapes. This cut-off happens for bodies larger than
a few km.
In Fig. 13 (right), the variations along the x-axis (e.g.
of the cut-off position) are due to different impact velocities
(for instance a larger planetesimal can deliver material onto
planet h because impacts happen at lower velocities) and it
can also vary with the planets' properties through H and
the atmospheric density (assumed constant for now) with
the scalings given by Eq. 9. Planets g and h can therefore
get volatiles delivered from larger comets than further in
planets. The variations along the y-axis are mainly due to
the fraction of comets accreted onto the planets and the
different impact velocities and scale as depicted by Eq. 11.
The effect of increasing Q from 1 to 100au (solid to
dashed lines) or increasing q from 10−5 to 10−1au/yr is the
same as explained when describing Fig. 13 (left), i.e., due to
the change in facc. This plot is therefore also general and can
be used to compute the outcome of an impact for any values
of q and Q, and is not tied to any of the specific scenarios
explained in Sec. 4.
The volatile mass that ends up in the atmospheres is a
fraction fvol of the mass delivered. We assume that volatiles
are delivered to the atmospheres in proportion to their frac-
tion of the mass of the parent body. For a comet-like body,
we assume a rock-to-ice mass ratio of 4 based on recent mea-
surements in the 67P comet (Rotundi et al. 2015), i.e 20%
c(cid:13) 2002 RAS, MNRAS 000, 1–28
5.3 The integrated effect of these impacts over
the age of the system
5.3.1 Total incoming mass over the system's age
We now work out the effect of impacts on the TRAPPIST-
1 planets over the age of the system and more specifically,
how much atmospheric mass is lost and how much projec-
tile/volatile mass is accreted for a given total incoming mass
of comets. To do so, we assume a typical N (D) ∝ Dγ size
distribution with γ = −3.5 for the comets that are expelled
from the belt (e.g. Dohnanyi 1969; Th´ebault & Augereau
2007) up to a maximum size of 10km8. Indeed, integrating
over the assumed size distribution for the total atmospheric
mass loss (or accreted material) shows that > 10km im-
pactors are unimportant (as already concluded by Schlicht-
ing et al. 2015) because Matmloss ∝ D−2 for large bodies
as seen from Fig. 13 (left), which decreases faster than the
gain in mass of these larger bodies (∝ D0.5). Very massive
giant impacts (e.g. Kral et al. 2015) of bodies with radius
> 1000km (i.e Pluto-sized or greater) can have a devastat-
ing effect on the atmosphere of a planet (Schlichting et al.
2015), which is not modelled in Shuvalov (2009), but these
impacts are rare and thus neglected in this study.
We consider an incoming mass of comets Minc that
reaches and can potentially hit the TRAPPIST-1 planets
after a mass Msca of comets has been scattered from this
outer belt over the system's age. Taking into account the ef-
ficiency to reach inner regions, Minc = Mscafin (see Fig. 14).
The integrated amount of mass scattered from a belt
Msca over the system's age can be evaluated. In Sec. 2, we
predicted that a potential planetesimal belt of 20M⊕ could
potentially have survived around TRAPPIST-1 at tens of
7 We assume that the bulk of the volatile mass is in water so
that this value is representative of the total volatile mass, though
a lower limit.
8 We note that the size distribution of the Kuiper belt for the
largest bodies is complicated and best-fitted by two shallow power
laws and a knee or a divot between the two (Lawler et al. 2018),
which would imply that most of the cross section would be in the
biggest bodies. This is not representative of what is observed in
general for the debris disc population, for which a -3.5 slope all the
way through the largest bodies is able to explain the observations.
16 Q. Kral et al.
Table 2. Table describing the parameters used for the different scenarios we tested. We list the rate of change of pericentre q, the
apocentre Q, the mean fraction accreted facc on each planet, the fraction of comets that makes it to the inner regions fin, the mass
fraction of volatiles on the exocomets/exoasteroids fvol, the minimum scattered mass Mscadestroy to destroy all 7 primordial atmospheres
(Msca = Minc/fin), the mass of delivered volatiles Mvolmin and water Mwatmin (assuming Solar-System comet-like compositions) for a
belt scattering at the low scattering rate of the current Kuiper belt (i.e Minc ∼ 10−2M⊕ fin) for each of the planets f, g, h, and MvolT,
MwatT for a belt of 20M⊕ (close to the expected mass for a potential leftover belt around TRAPPIST-1, see section 2) scattering 5%
(i.e Minc ∼ 1M⊕ fin) of its mass over 7Gyr. For the case of exoasteroids, fvol = 0.01, and Mvolmin as well as MvolT should be divided
by 20, and Mwatmin, MwatT by 10. Meo means 1 Earth ocean (i.e 2.5 × 10−4M⊕).
Scenarios
Single planet (1M⊕)
Single planet (10M⊕)
Planet chain (1M⊕)
Planet chain (10M⊕)
Kozai mechanism
q
au/yr
10−5
10−5
5 × 10−5
5 × 10−5
2 × 10−4
Q
au
1
1
1
1
100
facc
fin
fvol
Mscadestroy Mvolmin,Mwatmin
M⊕
M⊕,Meo
7 × 10−2
4 × 10−2
7 × 10−2
4 × 10−2
1.5 × 10−4
1
1
0.05
0.05
1
0.2/0.01
0.2/0.01
0.2/0.01
0.2/0.01
0.2
5 × 10−4
3 × 10−3
5 × 10−2
10−2
1
4 × 10−6,8 × 10−3
8 × 10−7,2 × 10−3
2 × 10−7,4 × 10−4
4 × 10−8,4 × 10−5
2 × 10−9,4 × 10−6
5.3.2 Atmospheric mass loss
MvolT,MwatT
M⊕,Meo
4 × 10−4,0.8
8 × 10−5,0.2
2 × 10−5,4 × 10−2
4 × 10−6,4 × 10−3
2 × 10−7,4 × 10−4
The total atmospheric mass loss for a given planet over the
system's age is
(cid:90) Dmax
Mtotatmloss =
N (D)Matmloss(D)facc dD,
(13)
Dmin
where N (D) is the number of bodies in each impactor di-
ameter bin D that make it to the inner regions.
Fig. 15 shows Mtotatmloss/Minc, i.e the total atmospheric
mass loss compared to the incoming mass Minc of comets
injected into the inner regions over the lifetime of the star.
Once again, this figure is general (and can be used for any q
and Q) and is not tied to a specific scenario (only the black
vertical lines are scenario dependent). We show the atmo-
spheric mass removed for specific values of apocentres Q = 1
and 100au but values for other Q can also be estimated (as
Mtotatmloss ∝ facc ∝ Q−1). Atmospheric mass loss remains
lower for planets g and h because impacts happen at lower
velocities (see also Fig. 13 left). The mean total atmospheric
mass loss for the seven planets can be approximated as
(cid:18)
(cid:19)−1(cid:18) Q
(cid:19)−1
Mtotatmloss
Minc
∼ 2 × 10
−3
q
10−5au/yr
,
(14)
1au
where we note that this ratio is accurate for planets d, e and
f but can be a factor 10 more or less for a specific planet (e.g.
10 times higher for planet b and 10 times lower for planet
h), and Fig. 15 should be used to get more accurate values.
To assess whether the impact process is capable of de-
stroying an entire primordial atmosphere, we first estimate
the primordial atmospheric masses of the different planets.
These primordial atmopsheric masses are not known and so
for reference we asssume an Earth-like composition and den-
sity. Computing the scale height for each planet (as in Eq. 9)
and assuming an isothermal atmosphere of temperature T
(the equilibrium temperature of the planets), we integrate
over the height of the planet atmospheres to get their masses
Matm = 4πρatm0H(R2
pla +2HRpla +2H 2). This gives primor-
dial atmospheric masses of 2, 0.9, 0.7, 0.9, 1.1, 0.7, 0.4 × 10−6
M⊕ for planets b to h. This is shown on Fig. 15 as horizontal
lines, where this mass has been divided by 1M⊕ to show the
effect of an incoming mass of 1M⊕.
Therefore, a primordial Earth-like density atmosphere
(12)
on the TRAPPIST-1 planets could be destroyed if
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Figure 14.
Schematic showing the main variables used to
parametrise comet scattering (here, for the planet chain case but
this is general) in the inner regions of the TRAPPIST-1 system.
au. By the action of a nearby planet, many planetesimals
may have been scattered inwards over the lifetime of the
system. Assuming that 5% of the belt mass is scattered over
7 Gyrs (using results by Marino et al. 2018), we get that
Msca ∼ 1M⊕ leading to Minc ∼ fin M⊕.
In our Solar System, ∼ 0.27 comet/yr leave the Kuiper
belt towards the inner regions (Levison & Duncan 1997).
The typical mass of comets in Levison & Duncan (1997)'s
study is ∼ 4 × 1013kg so that the rate of scattered incom-
Msca ∼ 2 × 10−3M⊕/Gyr. Therefore, a sim-
ing comets is
ilar Kuiper belt around TRAPPIST-1 would give Msca ∼
10−2M⊕ over 7Gyr leading to Minc ∼ 10−2fin M⊕.
However, the Kuiper belt is thought to have been a lot
more massive in its youth (e.g. Levison et al. 2011, and see
Sec. 2) and in general, debris discs that are observed can
have fractional luminosities of up to 104 greater than this
low-mass belt (Wyatt 2008), which is an indicator of them
being more massive. We note that the Kuiper belt is so light
(∼ 0.1M⊕, Fraser & Kavelaars 2009; Vitense et al. 2010)
that current instruments could not even detect it around
another star (Vitense et al. 2012; Kral et al. 2017b). From
an MMSN-like calculation, the initial Kuiper belt mass may
have been of several 10s of Earth masses (Hayashi 1981;
Morbidelli et al. 2003), meaning that Msca could have been
of the order of a few 10M⊕ owing to the depletion of the
belt to reach its current mass. In other words, we expect
−2fin M⊕ (cid:46) Minc (cid:46) 30fin M⊕.
10
PlanetscatteringQ=1au (apocentre)q<0.06au (pericentre) (pericentre evolution over time)Msca (comet mass scattered from belt)Minc= finMsca (comet mass that reaches inner region)Effects of impacts on the TRAPPIST-1 planets
17
Figure 15. Total atmospheric mass loss for a given incoming mass of comets that reach the inner region, i.e Mtotatmloss/Minc (assuming
a -3.5 size distribution) as a function of q for Q = 1 (solid lines) and 100au (dashed). The vertical black lines are typical
q values for
each scenario (see the text in Sec. 5.1). The primordial atmospheric masses of each planet are plotted as horizontal lines for an incoming
mass of 1M⊕, assuming an Earth-like atmospheric density and composition.
Minc > 5 × 10
−4
(cid:18)
(cid:19)(cid:18) Q
(cid:19)
1au
q
10−5au/yr
M⊕.
(15)
For the specific physical scenarios from Sec. 4 (see vertical
black lines on Fig. 15 for the planet scattering and Kozai sce-
narios), Table 2 shows the minimum scattered mass needed
Mscadestroy from an outer belt (the minimum incoming mass
would be finMscadestroy) to destroy the primordial atmo-
spheres of the seven planets.
For example, for the planet scattering scenario with
a single Earth-mass planet at 1au (i.e q ∼ 10−5au/yr,
Q = 1au and fin = 1, see Table 2), using Eq. 15 we find
that Minc (cid:38) 5 × 10−4M⊕ can destroy the primordial atmo-
spheres of the seven planets. This corresponds to a belt that
is being depleted for 7Gyr at a rate ten times lower than
that at which the current Kuiper belt is being depleted. If
the comets had to be passed in through a planetary sys-
tem before reaching the planet at 1au, the inefficiency in
the inward scattering process results in an additional fac-
tor fin = 0.05. This means that even with this factor, the
current Kuiper belt scattering rate is enough to destroy the
atmospheres of the seven TRAPPIST-1 planets.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
For the Kozai scenario q values are higher ( q ∼ 2 ×
10−4au/yr) and Q is at larger distances (100au), meaning
that interactions with planets are much more likely to re-
sult in ejections rather than accretions (see Fig. 3). We find
that Minc > 1M⊕ is needed to destroy the primordial atmo-
spheres, i.e two orders of magnitude larger than in the planet
chain case. For a 1M⊕ incoming mass (i.e. 100 times the cur-
rent Kuiper-belt like incoming mass rate), Fig. 15 shows that
the atmospheric mass loss is ∼ 2× 10−5M⊕ for planet b and
a factor 10 less for planets c, d, e, and f, and about another
factor 5-10 less for planets g, h (all of which are higher than
the primordial Earth-like atmospheric masses assumed here
except for planets g and h that are a factor 2 too small).
Given that the exo-Kuiper belts detected around F, G,
K stars are much more massive than the Kuiper belt, and
that the possible belt mass we derive for the TRAPPIST-
1 belt in Sec. 2 is ∼ 20M⊕), the scattering may be even
higher than assumed here (i.e., up to a factor of a few 103
the Kuiper-belt incoming mass), and we conclude that if a
scattering belt is around TRAPPIST-1, the primordial at-
mospheres would not survive impacts over the system's life-
time for both a planet scattering and a Kozai scenario.
10−610−510−410−310−210−1100q(au/yr)10−910−810−710−610−510−410−310−210−1Mtotatmloss/MincQ=100auQ=1auMatm/1M⊕PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth18 Q. Kral et al.
5.3.3 Water mass loss
In Table 3, we also quantify the maximum water mass loss
for the single and planet chain scenarios. The water mass
loss MwatLossT is given for each planet for a belt of 20M⊕,
which is close to the expected mass for a potential leftover
belt around TRAPPIST-1 (see section 2) scattering 5% (i.e.,
Minc ∼ 1M⊕ fin) of its mass over 7Gyr. For the planet chain
scenario, the planets can lose up to 4, 1.2, 0.8, 0.6, 0.4, 0.12,
0.06 Meo (Earth ocean mass), for b, c, d, e, f, g, h, respec-
tively, and 20 times more for the single planet case. These
values can be compared to the water mass loss from hy-
drodynamic escape due to XUV irradiation during the run-
away greenhouse phase, for which they found upper limits of
(Bourrier et al. 2017), 80, 40, 2.9, 1.5, 0.9, 0.4, 0.1 Meo, for
b, c, d, e, f, g, h, respectively. These values are, however, to
be taken as strict upper limits because it is uncertain that
hydrogen can reach the very top layers at the base of the
hydrodynamic wind, which is needed for it to escape (Bol-
mont et al. 2017). Also, this hydrodynamic escape works
well to eject hydrogen but other atoms are difficult to drag
along (Bolmont et al. 2017). For the impact case, not only
hydrogen would escape but the whole fluid in the ejected
plume. Bearing these caveats in mind, we can now compare
the water mass loss from hydrodynamic escape to the im-
pact scenario. For the planet chain case, the water mass loss
due to impacts seems to be less efficient than hydrodynamic
escape for planets b and c and both scenarios are within a
factor of a few for the other planets. For the most optimistic
case of the single planet case, impacts could produce the
same water loss as hydrodynamic escape for planets b and
c and be an order of magnitude higher for planets d to h.
5.3.4 Delivery of volatiles
We now evaluate the total mass of material and volatiles that
can be delivered from the impactors over the system's life-
time. We derive the total accreted projectile mass Mtotimpacc
by integrating the mass accreted per impactor (Fig. 13 right)
over the assumed size distribution. This accreted mass is as-
sumed to be deposited in the planets' atmospheres.
Fig. 16 shows Mtotimpacc/Minc, the total accreted pro-
jectile mass compared to mass of comets injected into the
inner regions over the lifetime of the star. The overall shape
is similar to Fig. 15, but note that planet h is far better
for delivery of mass into its atmosphere than having its at-
mosphere depleted because impacts are at lower velocities
(which means material from larger planetesimals can be ac-
creted, see Fig. 13 right). This means that the mass delivered
on planet h (and planets with similar impact velocities) may
be greater than that lost after each impact.
To quantify this, Fig. 17 shows the ratio of the accreted
projectile mass and atmospheric mass lost, which does not
significantly depend on q, instead only depending on the size
distribution of comets and slightly on Q. Thus, this ratio is
plotted as a function of the slope in the size distribution γ for
two different values of Q (1 and 100au). For γ = −3.5, this
ratio is greater than one for planets g and h and close to 1
for planet f but lower for the other planets. This means that
even if all of the accreted mass ends up in the atmosphere,
the total atmospheric mass must be decreasing for the inner
planets and can only increase for the outer three planets if
all mass ends up in the atmospheres. Regardless, all planets
will have their atmospheres enriched by the planetesimals'
composition and the situation is similar for all Q.
Consider now the fraction of this delivered projectile
mass that will be in volatiles, i.e Mtotvolacc = Mtotimpaccfvol,
which could be delivered to planets from comets. To assess
the amount of volatiles that are delivered to the planets
we consider two types of material that impact on to these
planets (presented in Sec. 5.2); 1) Cometary-like material
with 20% of ice by mass (fvol = 0.2), and 2) Asteroid-like
material with ∼1% of volatiles by mass (fvol = 0.01).
One important question is whether the icy material
will have dissappeared through sublimation before impact-
ing the planets. Marboeuf et al. (2016) show that a 1km
comet survives sublimation for ∼ 560 orbits around a 0.1L(cid:12)
star. TRAPPIST-1 is 200 times less luminous and so the
comets will survive much longer. Extrapolating Marboeuf
et al. (2016)'s formula to TRAPPIST-1 luminosity, we get
that a 1km comet passing at small pericentres (0.1au) would
need more than 105 orbits to sublimate. As it only takes a
few 100s of orbits for the comets to be accreted on the plan-
ets (see Fig. 4), we assume that most of the icy content of
the comets will not have sublimated and so will be available
to be delivered at impact. We note that during the q evolu-
tion, the sublimation will start happening only in the very
last phase, i.e when the pericentre is close to the planets
already (because for larger pericentres the mass loss from
comet sublimation is very slow and the timescale of evolu-
tion of the pericentre is much faster, Marboeuf et al. 2016).
Thus, the impact timescale of 100s of orbits is a good in-
dicator of the number of orbits before impact during which
sublimation could happen.
Therefore, we estimate the mean of the total volatile
mass delivered on each of the seven planets as
(cid:18)
(cid:19)−1(cid:18) Q
(cid:19)−1
,
1au
Mtotvolacc
Minc
∼ 2 × 10
−3fvol
q
10−5au/yr
(16)
which is, for all planets, within a factor 3 of that from
Fig. 16. For all planets, we can also estimate the incoming
mass needed to deliver more volatiles than the primordial
atmospheric mass
(cid:18)
(cid:19)(cid:18) Q
(cid:19)
1au
Minc > 5 × 10
−4f
−1
vol
q
10−5au/yr
M⊕,
(17)
where we assumed primordial atmospheres of Earth-like den-
sities.
Thus, for the planet scattering scenario, we find that
only a small incoming mass is needed to deliver enough
volatiles to potentially replenish an atmosphere with an
Earth-like density (e.g., Minc > 3 × 10−3M⊕ for comet-like
bodies scattered from an outer belt to a 1M⊕ planet at 1 au).
The incoming mass needed for the Kozai scenario is larger,
5M⊕, but not implausible to reach as shown by Eq. 12.
From Fig. 17, we have shown that only planets g and
h (and possibly f) would be able to retain the largest part
of the delivered volatiles. This means that for the planet-
scattering and Kozai scenarios, the new atmospheric compo-
sitions of planets f, g and h could be entirely set by the comet
volatile content, which would replenish the atmospheres over
the system's age. However, the absolute level of the volatile
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Table 3. Amount of water lost due to impacts for the planet scenario (single and chain). The water mass loss MwatLossT is given for
each planet for a belt of 20M⊕ (close to the expected mass for a potential leftover belt around TRAPPIST-1, see section 2) scattering
5% (i.e Minc ∼ 1M⊕ fin) of its mass over 7Gyr. Meo means 1 Earth ocean (i.e 2.5 × 10−4M⊕).
Effects of impacts on the TRAPPIST-1 planets
19
Scenarios
q
(au/yr)
Q
(au)
MwatLossT
(Meo)
Single planet (10M⊕)
Planet chain (10M⊕)
5 × 10−5
5 × 10−5
1
1
b
80
4
c
24
1.2
d
16
0.8
e
12
0.6
f
8
0.4
g
h
2.4
0.12
0.12
0.06
Figure 16. Total accreted projectile mass for a given incoming mass of comets sent to inner regions, i.e Mtotimpacc/Minc (assuming a
-3.5 size distribution) as a function of q for Q = 1 (solid lines) and 100au (dashed). The vertical black lines are typical q values for each
scenario (see the text in Sec. 5.1). The primordial atmospheric masses of each planet are plotted as horizontal lines for an incoming mass
of 1M⊕, assuming an Earth-like density and composition.
content that will remain in the atmosphere is difficult to con-
strain as some fraction of the volatile mass will be ejected
by later impacts or end up on the planet's surface and some
other sources of volatiles could be present (see Sec. 6.2). We
can, however, estimate the amount of volatiles that will sur-
vive after each impact assuming that a fraction fr of the
accreted material remains in the atmosphere rather than
condensing on the planet. Therefore, after a given impact
frMtotimpacc of material will be added to the atmosphere
and the next impact could remove a maximum of Mtotatmloss
from this added material. Assuming that fr = 1 (if impacts
are frequent enough, e.g. LHB-like, material does not have
c(cid:13) 2002 RAS, MNRAS 000, 1–28
time to condense back on the surface), we compute the frac-
tion of volatiles that would accumulate from subsequent im-
pacts in Fig. 18. We note that some additional volatiles could
be added by degassing of the planets' interiors but that fr
may also be smaller so that the exact volatile mass that
can accumulate depends on complex physics that cannot be
modelled in this paper. We see that indeed, only planets
f, g, and h have positive values (i.e. they gain volatiles over
time) and therefore appear9 in Fig. 18 showing Mvol/Mincvol,
9 However, for fr < 0.2, the atmospheric mass loss takes over
10−610−510−410−310−210−1100q(au/yr)10−910−810−710−610−510−410−310−210−1Mtotimpacc/MincQ=100auQ=1auMatm/1M⊕PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth20 Q. Kral et al.
and h might be more massive than that of the innermost
planets of the TRAPPIST-1 system if cometary bombarde-
ment has happened, and that a fraction of their composition
should reflect the cometary abundances in this system. We
note that the build-up of secondary atmospheres for planets
f, g and h is mainly allowed by the impact velocities that
are low enough on these outermost planets to both reduce
the atmospheric mass loss after each impact and allow to
deliver more volatiles (from larger bodies).
5.3.5 Delivery of water
Figure 17. Ratio of accreted projectile mass over atmospheric
mass lost for Q = 1 (solid) and 100au (dashed) for the different
planets as a function of the slope of the size distribution.
Mwater ∼ 2 × 10
−4
(cid:18) fvol
(cid:19)(cid:18) fwat
0.2
0.5
(cid:19)−1
Water on Solar System comets makes up more than fwat =
50% of the volatiles (Mumma & Charnley 2011). Depending
on fwat for exocomets, the amount of water Mwater delivered
on the seven planets can be approximated by
(cid:19)(cid:18)
(cid:18) Q
10−5au/yr
q
(cid:19)−1
1au
Minc,
(18)
where fvol ∼ 0.2 for exocomets (∼0.01 for asteroids) and
fwat ∼ 0.5 (∼1 for asteroids). For example, for the single
Earth-mass planet scattering scenario (i.e q ∼ 10−5au/yr,
Q = 1au, and fin = 1), we find that a belt scattering at
the same low rate as the current Kuiper-belt would result
in the planets accreting ∼ 8 × 10−3 Earth oceans of water
(or 10 times less for asteroid-like bodies), assuming that one
Earth ocean equals 1.5×1021kg (see Mwatmin in Table 2). We
note that for the planet chain case (where fin = 0.05), these
values would be a factor 20 smaller and for a larger incoming
mass Minc these values could go up by a factor more than
103 (see Eq. 12). We find that a belt of 20M⊕ (similar to
the plausible belt mass we predict around TRAPPIST-1 in
Sec. 2) that would scatter 5% of its mass over 7Gyr (i.e.,
Minc ∼ 1M⊕ fin) would deliver ∼ 1 Earth ocean of water
to the planets for the single planet case and ∼ 0.04 Earth
ocean for a planet chain (see MwatT in Table 2).
For the Kozai scenario, we find that between ∼ 10−5
(pessimistic case with a Kuiper belt scattering rate) and
∼ 0.01 (optimist case with Minc ∼ 20M⊕) Earth oceans of
water could be delivered to the planets.
This delivered water will presumably recondense as ice
on the surface of planet h (but when the star was younger
this planet was in the HZ and water could have been in
liquid form for a long period, see Sec. 6.1.2), but for warmer
planets such as planets f and g, we expect that a rain cycle
would create liquid water on these planets that would then
be reinjected into the atmospheres cyclically (see Sec. 6.3).
The temporal evolution of the build up of the amount of
water in these secondary atmospheres can be obtained from
Fig. 18 or from the coming Eq. 21 for planets f, g, and h.
6 DISCUSSION
Figure 18. Volatile mass that can accumulate Mvol in the atmo-
sphere impact after impact for a given incoming mass of cometary
volatiles Mincvol with Mincvol = Mincfvol as a function of q for
Q = 1 (solid lines) and 100au (dashed). The vertical black lines
are typical
q values for each scenario (see the text in Sec. 5.1).
The primordial atmospheric masses of each planet are plotted as
horizontal lines for an incoming mass in volatiles of 0.2M⊕ (i.e.
Minc = 1M⊕ for fvol = 0.2), assuming an Earth-like density and
composition.
where Mincvol = Mincfvol is the incoming mass of volatiles.
We can also derive a general formula as a function of q and
Q (similar to Eq. 16) that gives the mass of volatiles that
can accumulate Mvol rather than the total volatile mass de-
livered. We do that in Sec. 6.1.2 and give the temporal evo-
lution (assuming a constant rate of impact) of the build up
of the secondary atmospheres that are created for planets f,
g, and h (see Eq. 21).
We thus conclude that the atmospheres of planets f, g
for planets g and h (and for fr < 0.8 for planet f) so that no
secondary atmospheres would accumulate in this case, but this
neglects outgassing which would add more volatiles and would
make it harder to not build up secondary atmospheres on these
three planets.
6.1 Comparison between timescales of the
different processes
The consideration of timescales is important because it con-
strains the duration over which atmosphere loss/gain occurs
c(cid:13) 2002 RAS, MNRAS 000, 1–28
−4.0−3.8−3.6−3.4−3.2−3.0Slopeofsizedistributionγ0246810Mtotimpacc/MtotatmlossQ=1auQ=100auPlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth10−610−510−410−310−210−1100q(au/yr)10−910−810−710−610−510−410−310−210−1Mvol/MincvolQ=100auQ=1auMatm/0.2M⊕PlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlanethEffects of impacts on the TRAPPIST-1 planets
21
This shows that the timescales over which the primordial
atmospheres can be destroyed are much shorter than the age
of the system. Therefore, we confirm the previous conclusion
(see Sec. 5.3.2) that cometary impacts may have entirely
stripped all planets of their primordial atmospheres by 7Gyr,
even if the scattering rate is smaller by a factor more than 10
than assumed here (i.e., close to the Kuiper-belt scattering
rate level).
6.1.2 Timescale to regenerate secondary atmospheres
from impacts for planets f, g, and h
We also compute the temporal evolution of the volatiles Mvol
that are deposited and accumulate after each impact (i.e we
take into account that subsequent impacts remove part of
the volatiles delivered by the preceding impact as in Fig. 18).
For planets g and h, Mvol is given by (for any q and Q)
and is a factor 5 smaller for planet f. The amount of wa-
ter delivered is simply Mwater = fwatMvol. Now, we work
out the timescale to replenish the secondary atmospheres
of planets g and h in cometary volatiles at the level of a
10−6M⊕ atmospheric mass
treplenish ∼ 7Myr f
−1
in f
−1
vol
10−5au/yr
1au
(cid:18)
(cid:19)(cid:18) Q
(cid:19)
(cid:19)−1
q
(cid:18)
Msca
0.1M⊕/Gyr
,
(22)
and a factor 5 longer for planet f. The replenishment
timescale shows that in most physically motivated cases
planets f, g, and h will have had time (over the age of the
system) to rebuild secondary atmospheres with masses of at
least 10−6M⊕, i.e., equal or greater than an Earth-like pri-
mordial atmosphere. We note that most of the volatiles de-
livered by the comets have low condensation temperatures
and thus would remain in the atmosphere rather than go
on the planet's surface but water could condense as ice on
planet h and cycle from the surface to the atmosphere on
planets f, g owing to rain (see Sec. 6.3). Therefore, we ex-
pect Mvol to be a good estimate of the amounts of volatiles
that can accumulate for planets f and g and note that up to
50% of the volatiles (to account for water) could transform
into ice on planet g and thus reduce Mvol by a factor 2 (but
this ice could outgas at a later stage because of the planet
activity).
Fig. 20 shows an example for the planet chain scenario
with q = 10−5au/yr, Q = 1au, fin = 0.05 and fvol = 0.2
assuming Msca = 0.1M⊕/Gyr. We see that only planets f, g
and h can accumulate volatiles over time (see section 5.3.4).
For this planet chain scenario, planets g and h can get their
atmospheres replenished by cometary volatiles at the level of
their primordial atmospheres over a timescale of 500-700Myr
and ∼6Gyr for planet f.
We now compare the replenishment timescale to the
evolution of the position of the HZ. Luger & Barnes (2015)
Mvol(t) ∼ 10
−3finfvol
(cid:18)
(cid:18) t
10−5au/yr
q
(cid:19)(cid:18)
(cid:19)−1(cid:18) Q
1au
(cid:19)−1
(cid:19)
Msca
7Gyr
0.1M⊕/Gyr
M⊕,
(21)
Figure 19. Atmospheric mass loss Matmlossc as a function of
time for a belt scattering at a constant rate Msca = 0.1M⊕/Gyr
for 7Gyr. We assume the planet chain scenario with q =
10−5au/yr, Q = 1au, fin = 0.05. The primordial atmospheric
mass of each planet is plotted as a horizontal line assuming an
Earth-like density.
compared with other processes which may be taking place,
but which are beyond the scope of this manuscript to con-
sider in detail.
6.1.1 Timescale to lose primordial atmospheres from
impacts
Assuming a constant rate of scattering Msca over 7Gyr, we
compute the atmospheric mass lost as a function of time
Matmlossc(t) ∼ 2 × 10
−3fin
(cid:18) t
q
10−5au/yr
Msca
(cid:18)
(cid:19)(cid:18)
(cid:19)−1
(cid:19)−1(cid:18) Q
(cid:19)
1au
7Gyr
0.1M⊕/Gyr
M⊕,
(19)
Msca = 0.1M⊕/Gyr corresponds to a
where we note that
belt with a total incoming mass of ∼1M⊕ fin, i.e similar to
what would be expected for a 20M⊕ belt scattering 5% of
its material over the age of the star. Matmlossc(t) becomes
greater than an atmospheric mass of 10−6M⊕ for
(cid:18)
tdestroy ∼ 4Myr f
−1
in
q
10−5au/yr
(cid:19)
(cid:19)(cid:18) Q
(cid:18)
1au
Msca
0.1M⊕/Gyr
(cid:19)−1
.
(20)
Now, we consider the planet chain scenario (i.e with
q = 10−5au/yr, Q = 1au, and fin = 0.05) with a scattering
rate Msca = 0.1M⊕/Gyr and look at the temporal evolution
of the atmospheric mass loss Matmlossc(t) due to the series of
impacts over the system's age as shown by Fig. 19. By com-
paring to the primordial atmospheric masses of the planets
(horizontal lines in Fig. 19), we see that for this scenario,
it takes between 10 and 400Myr to destroy the primordial
atmospheres of all seven planets (assuming an Earth-like at-
mospheric density). This is very fast compared to the age of
the system.
c(cid:13) 2002 RAS, MNRAS 000, 1–28
105106107108109Time(yr)10−810−710−610−510−410−3Matmlossc(M⊕)MatmPlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlaneth22 Q. Kral et al.
Figure 20. Volatile mass that accumulates impact after impact
Mvol as a function of time. We assume the planet chain scenario
with q = 10−5au/yr, Q = 1au, fin = 0.05 and fvol = 0.2 and
Msca = 0.1M⊕/Gyr.
a constant comet input rate over 7Gyr of
The primordial atmospheric mass of each planet is plotted as a
horizontal line assuming an Earth-like density.
show that for a 0.08M(cid:12) star, the HZ location moves inwards
to its present-day position after ∼ 1Gyr. This means that
planet h will be the first to enter the liquid water HZ, which
it will do at a point when the closer-in planets are still in
a runaway greenhouse state (assuming they have retained
any atmospheres). According to the Luger & Barnes (2015)
model, planet h crosses into the empirical habitable zone at
∼30Myr. Coupled with our results, this scenario indicates
that planet h could have received significant volatile deliv-
ery at a point in its history (i.e., between 30Myr and 1Gyr)
when liquid water was stable at its surface (Fig. 20). This
raises the prospect for an early carbon cycle being estab-
lished on this planet, stabilising climate through water-rock
interaction as is inferred for Earth (Walker et al. 1981).
6.2 Additional sources of volatiles
6.2.1 Volatiles created by vapourised material from the
planet's surface during impact
The volatile fraction that ends up in the atmospheres of the
TRAPPIST-1 planets does not only build up from the im-
pactor material but also from the vapourised material from
the planet surface, as was probably the case for the Chicx-
ulub impact that may have released large quantities of gas
and dust contributing to the environmental stress that led
to the demise of dinosaurs on Earth (Pope et al. 1997).
From Okeefe & Ahrens (1977), we can estimate the vol-
ume of material Vvap vapourised from a given meteoritic
impact (with a volume Vpr). They find that Vvap = 0.4SVpr,
where S = (ρpr/ρt)(Vimp/Cp)2, using the same notations
as in previous sections and Cp being the bulk sound speed
of planetary surface, which varies depending on the planet
ground composition (Melosh 1989). We assume an Earth-
like composition for which Cp ∼ 7km/s. We thus find that
the vapour mass Mvap produced for a given impactor of mass
Mimp is Mvap = 0.4Mimp(Vimp/(7km/s))2.
However, some of the vapour ejecta will escape and only
a fraction will have a low enough velocity to be retained in
the atmosphere. Once again, using results from Shuvalov
(2009), we get that the maximum ejected fraction of target
material after impact is Mtaresc ∼ 0.02Mimp(Vimp/Vesc)2.
This maximum is reached for bodies that are larger than
∼1km and for smaller bodies, the planet retains almost all
of the target material created at impact. Of course, above
a certain threshold it means that the whole target mass es-
capes (as Mtaresc becomes greater than the total atmospheric
mass), which is similar to the projectile mass behaviour
(where volatiles from bodies larger than ∼10km cannot be
retained in the atmosphere). We thus find that Mvap is a
good indicator of the vapourised mass that will remain in
the atmosphere (as Mtaresc (cid:28) Mvap).
From Eq. 11, we notice that the mass delivered from
the projectile quickly tends to Mimp for bodies smaller than
about 10km. Thus, for planets g and h that have median
impact velocities of ∼ 25 and 20km/s, Mvap will be slightly
higher but of the same order of magnitude as Mimpacc. This
means that some volatiles such as SO2, CO2 or water could
also be formed from the vapourised planets' crust (see Pope
et al. 1997). However, we note that the typically low con-
centration of volatiles in planetary basalts that would form
the bulk of a crust would not release as many volatiles as
for the Chicxulub impact (e.g. Dreibus & Wanke 1987; Saal
et al. 2002).
6.2.2 Outgassing on the planets
Degassing may happen early during accretion when form-
ing the planets but this is not a concern in our study as we
expect the primordial atmospheres to be totally destroyed.
Degassing from tectonic activity may also happen at a later
stage that could affect the amount of volatiles in the atmo-
spheres. Another way of producing degassing is from stellar
induction heating. A recent study that focused on the effect
of this mechanism on the TRAPPIST-1 planets finds that
induction heating could create strong outgassing on planets
b, c, d that are very close to their host star but it should
not affect the outermost planets e, f, g, and h (Kislyakova
et al. 2017).
For the plate-tectonic degassing, we take the degassing
on Earth as an upper bound because plate tectonics is very
active on Earth and may be less efficient/active on other
planets10. Earth produces ∼ 22km3 of basaltic magmas each
year (Crisp 1984). Given a magma density of 2600kg/m3, we
estimate a total degassing rate of ∼ 6×1013kg/yr. Assuming
a typical water content of 0.3wt% and the extreme case of
perfectly efficient degassing with no subduction recycling of
water to the planet's mantle, we find that an upper bound
on the tectonically driven water degassing rate is ∼ 3 ×
10−5 M⊕/Gyr (0.11 Earth oceans per Gyr). Therefore, if
the tectonic activity on planets f, g and h were as active as
on Earth, degassing of water could occur at a similar rate
to the water delivered from impacting comets (see Table 2),
thus enhancing the amount of water on planets f, g and h.
10 We note that a recent study shows that even for planets in
the stagnant lid regime (i.e., without plate-tectonics), volcanic
outgassing rates suitable for habitability could possibly be main-
tained (Foley & Smye 2017).
c(cid:13) 2002 RAS, MNRAS 000, 1–28
107108109Time(yr)10−810−710−610−510−410−3Mvol(M⊕)MatmPlanetbPlanetcPlanetdPlanetePlanetfPlanetgPlanethEffects of impacts on the TRAPPIST-1 planets
23
6.2.3 Volatiles that are ejected of the atmosphere and
reaccreted later
The material that escapes the planetary atmospheres after
each impact because they have velocities greater than the
escape velocity will end up in an eccentric torus around the
star close to the given planet location (e.g. Jackson et al.
2014; Cataldi et al. 2017). The eccentricity will vary depend-
ing on the ejection velocity of the material. While we expect
that high-velocity ejecta may reach neighbouring planets
(e.g. in a Panspermia-manner, Krijt et al. 2017; Lingam &
Loeb 2017), most of the material in the torus would interact
with the planet it has been ejected from. We note that for
an Earth-like planet on a slightly wider orbit than planet h,
the escape velocity (of about 10km/s) could become greater
than the planet's Keplerian velocity and thus the material
would not form a torus but rather be ejected on unbound
orbits.
The fate of the material in the torus is not straightfor-
ward to model. The material could deplete collisionally due
to high-velocity collisions in the elliptic torus and be ground
down to dust, which would be blown out from the system
by stellar wind radiation pressure (Wyatt 2008) and at the
same time eject the ices or volatiles present on the grains.
While the ejecta is also partly made up of gas, one could also
expect that the gas material (at least the fraction that is not
blown out by radiation pressure) in the torus will viscously
spread (maybe dragging dust with it) and end up on more
distant planets. The fate of the material that would be able
to interact with a planet for a long enough timescale is to be
reaccreted onto the progenitor planet (Wyatt et al. 2017).
The exact outcome depends on the exact chemico-physical
conditions in the TRAPPIST-1 planets environment, which
is not known, and thus goes beyond the scope of this paper.
6.3 Composition of the atmospheres at the end of
the impact process
Thanks to our model, we are able to retrieve the amount of
volatiles that is delivered to the different planets as well as
the atmospheric mass removed by a long-term series of im-
pacts. For the outermost planets, we find that the volatiles
delivered by impacts may accumulate and be abundant,
which could give us a way to constrain the atmospheric
composition of planets f, g, h, the former two being in the
HZ. However, we need to understand how these delivered
volatiles would evolve in their new atmospheres to predict
the current atmospheric compositions of these planets. They
could chemically react to form new species, condense on the
surface as ice and some additional volatiles may be produced
as seen in the previous Sec. 6.2.
For instance, the delivered water will presumably con-
dense as ice on the surface of the colder planet h (when it
finishes being in the HZ, see Sec. 6.1.2) but for warmer plan-
ets in the HZ (e.g. planets f and g), a rain cycle could create
liquid water on the planets that is then reinjected into the
atmospheres cyclically. Volatiles such as CO, CO2, or CH4
have a low condensation temperature and will remain in the
atmosphere along with other similar volatiles delivered by
the comets. However, when liquid water is on the planet,
this can draw the CO2 content down by silicate weathering
c(cid:13) 2002 RAS, MNRAS 000, 1–28
that fixes CO2 in the planet's surface (forming carbonates)
as shown in Siever (1968).
Over longer timescales, these volatiles can further chem-
ically react to form new molecules. However, the exact com-
position of the delivered volatiles depends on the composi-
tion of exocomets in our scenario. The latter has been found
to be consistent with the composition of comets in our Solar
System (e.g. Matr`a et al. 2017a) but there is still a wide
range of observed compositions amongst the Solar System's
comets (e.g. Mumma & Charnley 2011).
Another complication is that, as discussed in the pre-
vious subsection 6.2, volatiles may also be formed from the
vapourised planet's crust during impact, from outgassing
and even by reaccretion of previously ejected material, which
would mix with the volatiles delivered by impacts.
All of these factors (active chemistry, potential addi-
tional volatiles, exocomet composition) makes it hard to pre-
dict the exact final compositions of the atmospheres after a
few Gyr of evolution. An atmosphere model that would make
assumptions about what happens without impacts could be
fed by our impact predictions to come up with a plausible
likely composition, but this goes beyond the scope of the
present paper. We note however that these extra sources of
volatiles do not change our conclusion that in the presence of
a belt scattering comets, the atmospheres of the outermost
planets f, g, h should be more massive.
6.4 Impacts in very dense Venus-like atmospheres
We note that our model is not valid for very massive atmo-
spheres. If the atmospheres of planets f, g, h become mas-
sive enough (Venus-like, i.e., 200bars) due to impacts (or
if the primordial atmospheres were Venus-like), 1-10km im-
pactors do not create craters anymore but rather decelerate
and get fragmented before touching the ground and create
big aerial bursts that are very effective at removing atmo-
spheric mass (Shuvalov et al. 2014). The amount of accreted
projectile material is also very high (close to 100%) for these
aerial burst type of impacts (Shuvalov et al. 2014). There-
fore, for very dense atmospheres, we expect an increased
delivery of volatiles from the impactors and less from the
vapourised crust. We also expect that Venus-like primordial
atmospheres would still be destroyed, since those impacts
are more effective at removing mass, therefore not changing
our conclusions.
6.5 Implications for life on these planets
One of the prime motives in searching for planets orbiting
very low mass stars is to study the chemical composition of
their atmospheres, and discover whether they contain large
quantities of gas of a likely biological origin (e.g. Seager et
al. 2016). Here, we consider the implications of our results
concerning impacts towards creating the first forms of life.
Many elements can affect the emergence of life, most of
which currently remain unconstrained empirically. We chose
to apply our study to the TRAPPIST-1 system because its
seven planets mark an important milestone. In addition to
the multiple advantages of having a very low-mass host star
for atmospheric characterisation (e.g. He et al. 2017), these
seven worlds allow us to compare each to one another. All
24 Q. Kral et al.
seven have followed a similar history in terms of UV irradi-
ation for instance (modulo their distance to the star). Here
we have tried to quantify whether all planets would receive
a similar impact history, which may be important to kick
start life as explained further.
UV irradiation has often been seen as prejudicial to hab-
itability. Its main disadvantages are: 1) to photodissociate
water molecules, of which the hydrogen is then lost to the
space, depleting its oceans (e.g. Bourrier et al. 2017), and
2) to break complex molecules on the surface, and affect
replication (e.g. O'Malley-James & Kaltenegger 2017). The
situation is particularly sensitive for planets orbiting very
low-mass stars like TRAPPIST-1, since these spend a long
time contracting onto the main-sequence, in a 1 Gyr stage
of particularly heightened far UV activity (e.g. Rugheimer
et al. 2015).
However, these issues might be mitigated by several
effects: 1) Ocean loss depends on the initial water reser-
voir (e.g. Ribas et al. 2016; Bolmont et al. 2017), and the
TRAPPIST-1 planets might have been initially rich in wa-
ter, having possibly assembled beyond the snow-line (Alib-
ert & Benz 2017; Ormel et al. 2017) and/or accreted water
at a later stage owing to impacts (as shown in this study);
2) UV photons do not penetrate water well, and organisms
can protect themselves under a few metres of water (e.g. Es-
trela & Valio 2017); 3) UV irradiation accelerates mutations,
leading to Darwinian evolution; 4) the non-illuminated side
of a tidally synchronised planet is protected; and 5) UV ir-
radiation, impacts, and a hard surface might be required to
kick-start life (abiogenesis).
The literature contains much debate on many of the
points above, except on the very last one, which we describe
in more detail here as it is related to the outcome of this
paper.
Recent advances
in biochemistry (summarised in
Sutherland 2017) have shown a prebiotic chemical path lead-
ing from hydrogen cyanide (HCN) to formaldehyde (CH2O),
a known precursor to ribonucleotides (the building block
to biologically relevant molecules such as ATP, RNA and
DNA), amino acids (required for proteins) and lipids (Patel
et al. 2015). Hydrogen cyanide, the initial molecule needed
to inititate the process, can be produced in the plasma cre-
ated when impactors enter in contact with an atmosphere
(Ferus et al. 2015). In the presence of UV radiation, hy-
drogen cyanide can then react with other compounds that
can be found concentrated on a planetary surface to cre-
ate the building blocks of life. The impactor itself may have
another role to play, which is to excavate underground ma-
terial, and reveal chemically interesting strata (Patel et al.
2015), thereby acting as a chemical reactor.
We show in this paper that, if a belt scattering comets is
present in the system, numerous impacts with different ener-
gies will happen throughout the history of the TRAPPIST-1
planets. From these impacts, we expect to create a subse-
quent amount of HCN in the impactor plasma (Ferus et al.
2015). We also note that as HCN is found in comets (e.g.
Mumma & Charnley 2011), it may also be present on the
potential exocomets of TRAPPIST-1 and be delivered along
with the other volatiles (e.g. see Matr`a et al. 2017b). We also
emphasise that if the planets are tidally locked, it does not
affect the emergence of life in this scenario as we predict that
about half of the impacts would happen on the night side
and the other half on the day side so that the UV photons
from the star necessary for reactions to happen will be able
to play their role.
Thus, our scenario offers the seed to create the first
building blocks of life and more detailed modelling is needed
to quantify how many ribonucleotides, amino acids and
lipids could be created from the impact properties (e.g. im-
pact velocities, rate of impacts) we predict. This is beyond
the scope of this paper but should give birth to new inter-
esting studies in the near future. Panspermia may also be
viable to transport some potential life forms to other plan-
ets, which can enhance the probability of life spreading in
the system (Krijt et al. 2017; Lingam & Loeb 2017).
To conclude, we cannot be certain yet that such a path
is where biology originated, however, it provides a different
narrative, one that requires UV irradiation, impacts and a
limited amount of water. Ultraviolet, in this context, be-
comes beneficial by removing excess liquid water and trans-
forming hydrogen cyanide into formaldehyde, whereas im-
pacts would bring in energy to create hydrogen cyanide,
and replenish the planet in volatiles such as water, much
like what happened in the LHB (e.g. Court & Sephton 2014;
Nesvorn´y et al. 2017) after a desiccating moon-forming im-
pact (e.g. Canup 2014).
7 CONCLUSION
In this paper, we have studied the effects of impacts on the
seven TRAPPIST-1 planets in terms of atmospheric mass
loss and delivery of volatiles and water. We derive general
results for any scenario where the comet pericentres slowly
migrate inwards at a rate q. We also specifically test three
scenarios for the delivery of comets from an outer belt to the
inner planets (located within 0.1au): 1) Planet scattering by
a single or a chain of planets, 2) An outer companion forcing
Kozai oscillations on comets leading them to small pericen-
tres, 3) Galactic tides on an exo-Oort cloud. We model these
three scenarios by a steadily decreasing pericentre (constant
q) that is quantified in Sec. 4 for each of the scenarios. The
results can be summed up as follows:
• We find that applying a minimum mass TRAPPIST-
1 nebula approach lead to a surface density Σ ∼
122 (r/1au)−1.97 kg/m2. We show that a potential belt
around TRAPPIST-1 could not survive within 10au because
of collisional erosion (if it was created at the end of the proto-
planetary disc phase). Assuming that such a belt is between
10 and 50au, and extrapolating the derived minimum sur-
face density, we infer that this belt would have mass of at
least 20M⊕ and may be observable in the far-IR or sub-mm
with ALMA (see Sec. 2).
• We ran a suite of N-body simulations to understand
the dynamics of comets that impact onto the seven different
planets. We find the impact and ejection probabilities for
each comet's orbit (see Figs. 2 and 3). We also provide the
accretion timescales for these different comet families (see
Fig. 4). We analytically explain the main dependencies for
these probabilities and timescales.
• We give the impact velocity distributions for each
planet, and we find that they typically have double-peaked
profiles (see Fig. 6). The median impact velocity for planet b
is close to 100km/s, whilst for planet h, it is close to 20km/s
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Effects of impacts on the TRAPPIST-1 planets
25
(see Fig. 7). These impact velocities are always much above
the escape velocities of the planets and gravitational focus-
ing is not important.
• We find that the fraction of comets accreted on each
planet depends on the decreasing rate of pericentres ( q) and
apocentre Q (scaling as q−1Q−1). We find two regimes, for
small q, most of the impacts end up on planets g and h and
for higher q, each planet gets a fraction of comets accreted
(see Fig. 12).
• The atmospheric removal is dominated by comets of a
few km in diameter (see Fig. 13 left).
• The delivery of volatiles is only possible for comets
(cid:46) 3km in size (see Fig 13 right). For bigger comets, the
projectile material escapes and no delivery is possible.
• We find that the higher impact velocities for the inner-
most planets lead to a higher atmospheric removal rate for
a given cometary impact rate and a lower amount of volatile
delivered.
• In general, we find that if the incoming mass of
regions Minc > 5 ×
(cid:1) M⊕, the primordial atmospheres of
10−4(cid:16)
(cid:17)(cid:0) Q
reach the inner
comets
that
q
10−5au/yr
1au
q
−1
vol
(cid:16)
10−5au/yr
(cid:17)(cid:0) Q
(cid:1) M⊕ (where fvol is
the seven planets would be totally destroyed (see Fig. 15),
i.e a belt with a low scattering rate similar to the current
Kuiper belt is enough to destroy all primordial planetary at-
mospheres.
• We quantify the amount of water lost owing to impacts
and find that it is similar (possibly higher) to the amount of
water lost through hydrodynamic escape (see Sec. 5.3.3 and
Table. 3).
• As for the delivery of volatiles to the comets (see
Fig. 16), we find that planets g and h (and most likely f)
may retain volatiles from the impacting comets in their at-
mospheres and the conclusion holds for any size distribution
of incoming comets between -3 and -4 (see Fig. 17).
• We thus predict that if the planets were hit by comets,
the atmospheres of planets f, g, and h would be more massive,
which could be checked by future missions in the next decade.
• We also show that for an incoming mass of comets
Minc > 5 × 10−4f
the volatile fraction on solids), the volatile mass delivered
by comets is greater than Earth-like atmospheric masses (as-
suming Earth-like densities for the 7 planets).
• We provide a prescription for the amount of water or
volatiles that can accumulate as a function of time (see
Eq. 21) that could be used to feed an atmospheric model
to check the actual composition of atmospheres dominated
by the delivery of comets.
• We find that a large quantity of volatiles may have been
delivered to planet h while it was still in the liquid water
habitable zone.
• We find that a planet chain that would scatter comets
from an exo-Kuiper belt or an outer companion that would
force Kozai oscillations on a comet belt are two plausible
mechanisms to throw an important number of comets on
the seven planets over the system's lifetime (see Secs. 4.1
and 4.2).
• On the other hand, we rule out a potential Oort-cloud
around TRAPPIST-1 as being a significant source of im-
pacting comets (see Sec. 5.1.2).
• For the planet-scattering scenario, we find that even a
belt with a low scattering rate similar to the current Kuiper-
1au
c(cid:13) 2002 RAS, MNRAS 000, 1–28
belt is enough to destroy typical Earth-like primordial at-
mospheres for the seven planets. Taking into account that
typically observed debris belts are much more massive than
the Kuiper belt, we find that the Kozai (slightly less effi-
cient) scenario can also strip primordial atmospheres even if
the impact process only lasts a fraction of the system's age.
• As for the volatile delivery, we find that for the planet-
scattering scenario, planets f, g, and h can get (more than)
an Earth ocean mass of water (and other volatiles) delivered,
which can accumulate impact after impact. We find that the
primordial atmospheres are gradually replaced by cometary
material and may lead to subsequent build up of new sec-
ondary atmospheres with exocomet-like compositions. These
new secondary atmospheres may become more massive than
the initial primordial atmospheres.
• Table 2 summarises the results for the different scenar-
ios as for the minimum scattered (incoming) mass needed to
destroy the primordial atmospheres and the volatile/water
masses that can be delivered onto each planet.
• We also discuss the implications of impacts to create
the building blocks of life. We detail new emerging pathways
that can lead to life showing that UV irradiation, impacts
and a hard planetary surface might be enough to kick start
biological reactions and form ATP, RNA, DNA, amino acids
and lipids that are essential to life (see Sec. 6.5).
In brief, we find that the primordial atmospheres of the
seven planets orbiting around TRAPPIST-1 would not sur-
vive over the lifetime of the system if a belt scattering comets
at a similar low rate than the Kuiper belt (or faster) were
around TRAPPIST-1. According to our calculations based
on applying a minimum mass extrasolar nebula approach for
the TRAPPIST-1 system, we expect a potential 20M⊕ belt
may have survived around TRAPPIST-1 that would be ob-
servable with ALMA. We also show that a large fraction of
the delivered cometary volatiles remains in the atmospheres
of the outermost planets f, g and h, which gradually replace
their primordial atmospheres. We predict that the new sec-
ondary atmospheres of planets f, g and h may be more mas-
sive than that of the innermost planets (which may soon
be checkable/observable with the JWST) and their compo-
sition might be dominated by the composition of exocomets
in this system (i.e., impacts leave an imprint). We also pre-
dict that more than an Earth ocean mass of water could be
delivered to planets f, g, and h owing to impacts that may
be in liquid form on planets f and g.
ACKNOWLEDGMENTS
This paper is dedicated to Mila. We thank the two refer-
ees for comments that greatly improved the quality of the
paper. QK and MCW acknowledge funding from STFC via
the Institute of Astronomy, Cambridge Consolidated Grant.
QK thanks J. Teyssandier for interesting discussions about
the Kozai mechanism. Simulations in this paper made use
of the REBOUND code which can be downloaded freely at
http://github.com/hannorein/rebound.
26 Q. Kral et al.
REFERENCES
Abe, Y., Ohtani, E., Okuchi, T., Righter, K., & Drake, M.
2000, Origin of the Earth and Moon, 413
Alberti, T., Carbone, V., Lepreti, F., & Vecchio, A. 2017,
arXiv:1706.06005
Alibert, Y., & Benz, W. 2017, A & A, 598, L5
Anglada, G., Amado, P. J., Ortiz, J. L., et al. 2017, ApJ,
850, L6
Antognini, J. M. O. 2015, MNRAS, 452, 3610
Baraffe, I., Homeier, D., Allard, F., & Chabrier, G. 2015,
A & A, 577, A42
Barstow, J. K., & Irwin, P. G. J. 2016, MNRAS, 461, L92
Batygin, K., & Morbidelli, A. 2017, AJ, 154, 229
Beust, H., Vidal-Madjar, A., Ferlet, R., & Lagrange-Henri,
A. M. 1990, A & A, 236, 202
Bolmont, E., Selsis, F., Owen, J. E., et al. 2017, MNRAS,
464, 3728
Bonsor, A., Augereau, J.-C., & Th´ebault, P. 2012, A & A,
548, A104
Bonsor, A., Raymond, S. N., Augereau, J.-C., & Ormel,
C. W. 2014, MNRAS, 441, 2380
Booth, M., Wyatt, M. C., Morbidelli, A., Moro-Mart´ın, A.,
& Levison, H. F. 2009, MNRAS, 399, 385
Boss, A. P., Weinberger, A. J., Keiser, S. A., et al. 2017,
arXiv:1708.02200
Bourrier, V., Ehrenreich, D., Wheatley, P. J., et al. 2017,
A & A, 599, L3
Breiter, S., & Ratajczak, R. 2005, MNRAS, 364, 1222
Burgasser, A. J., & Mamajek, E. E. 2017, arXiv:1706.02018
Canup, R. M. 2014, Philosophical Transactions of the Royal
Society of London Series A, 372, 20130175
Carone, L., Keppens, R., & Decin, L. 2016, MNRAS, 461,
1981
Cataldi, G., Brandeker, A., Th´ebault, P., et al. 2017, As-
trobiology, 17, 721
Chiang, E., & Laughlin, G. 2013, MNRAS, 431, 3444
Court, R. W., & Sephton, M. A. 2014, GCA, 145, 175
Crisp, J. A. 1984, Journal of Volcanology and Geothermal
Research, 20, 177
Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298
Ferus, M., Nesvorn´y, D., Sponer, J., et al. 2015, Proceed-
ings of the National Academy of Science, 112, 657
Foley, B. J., & Smye, A. J. 2017, arXiv:1712.03614
Fouchard, M., Froeschl´e, C., Valsecchi, G., & Rickman, H.
2006, Celestial Mechanics and Dynamical Astronomy, 95,
299
Fraser, W. C., & Kavelaars, J. J. 2009, AJ, 137, 72
Frewen, S. F. N., & Hansen, B. M. S. 2014, MNRAS, 439,
2442
Gaidos, E. 2017, MNRAS, 470, L1
Gillon, M., Jehin, E., Lederer, S. M., et al. 2016, Nature,
533, 221
Gillon, M., Triaud, A. H. M. J., Demory, B.-O., et al. 2017,
Nature, 542, 456
Godolt, M., Grenfell, J. L., Kitzmann, D., et al. 2016, A &
A, 592, A36
Grimm, S. L., Demory, B.-O., Gillon, M., et al. 2018,
arXiv:1802.01377
Haworth, T. J., Facchini, S., Clarke, C. J., & Mohanty, S.
2018, MNRAS, 475, 5460
Hayashi, C. 1981, Progress of Theoretical Physics Supple-
ment, 70, 35
He, M. Y., Triaud, A. H. M. J., & Gillon, M. 2017, MNRAS,
464, 2687
Heisler, J., & Tremaine, S. 1986, Icarus, 65, 13
Hendler, N. P., Mulders, G. D., Pascucci, I., et al. 2017,
ApJ, 841, 116
Howell, S. B., Everett, M. E., Horch, E. P., et al. 2016,
ApJ, 829, L2
Innanen, K. A., Zheng, J. Q., Mikkola, S., & Valtonen,
M. J. 1997, AJ, 113, 1915
Jackson, A. P., Wyatt, M. C., Bonsor, A., & Veras, D. 2014,
MNRAS, 440, 3757
Janson, M., Hormuth, F., Bergfors, C., et al. 2012, ApJ,
754, 44
Kaib, N. A., & Quinn, T. 2009, Science, 325, 1234
Kennedy, G. M., & Wyatt, M. C. 2010, MNRAS, 405, 1253
Kiefer, F., Lecavelier des Etangs, A., Boissier, J., et al.
2014, Nature, 514, 462
de Niem, D., Kuhrt, E., Morbidelli, A., & Motschmann, U.
Kislyakova, K. G., Noack, L., Johnstone, C. P., et al. 2017,
2012, Icarus, 221, 495
de Wit, J., Wakeford, H. R., Gillon, M., et al. 2016, Nature,
537, 69
arXiv:1710.08761
Kozai, Y. 1962, AJ, 67, 591
Kral, Q., Th´ebault, P., Augereau, J.-C., Boccaletti, A., &
de Wit, J., Wakeford, H. R., Lewis, N., et al. 2018,
Charnoz, S. 2015, A & A, 573, A39
arXiv:1802.02250
Kral, Q., Wyatt, M., Carswell, R. F., et al. 2016, MNRAS,
Di Sisto, R. P., Fern´andez, J. A., & Brunini, A. 2009,
461, 845
Icarus, 203, 140
Dohnanyi, J. S. 1969, J. Geophys. Res. , 74, 2531
Dreibus, G., & Wanke, H. 1987, Icarus, 71, 225
Duncan, M. J., & Levison, H. F. 1997, Science, 276, 1670
Duncan, M., Quinn, T., & Tremaine, S. 1987, AJ, 94, 1330
Ehlmann, B. L., Anderson, F. S., Andrews-Hanna, J., et
al. 2016, Journal of Geophysical Research (Planets), 121,
1927
Kral, Q., Krivov, A. V., Defr`ere, D., et al. 2017, The As-
tronomical Review, 13, 69
Kral, Q., Clarke, C., & Wyatt, M. 2017, Handbook of Ex-
oplanets, Edited by Hans J. Deeg and Juan Antonio Bel-
monte. Springer Living Reference Work, ISBN: 978-3-319-
30648-3, 2017, id.165, 165
Kral, Q., Matr`a, L., Wyatt, M. C., & Kennedy, G. M. 2017,
MNRAS, 469, 521
Eiroa, C., Rebollido, I., Montesinos, B., et al. 2016, A &
Krijt, S., Bowling, T. J., Lyons, R. J., & Ciesla, F. J. 2017,
A, 594, L1
ApJ, 839, L21
Eiroa, C., Marshall, J. P., Mora, A., et al. 2013, A & A,
Lawler, S. M., Shankman, C., Kavelaars, J. J., et al. 2018,
555, A11
AJ, 155, 197
Emel'Yanenko, V. V., Asher, D. J., & Bailey, M. E. 2007,
MNRAS, 381, 779
Levison, H. F., & Duncan, M. J. 1997, Icarus, 127, 13
Levison, H. F., Duncan, M. J., Brasser, R., & Kaufmann,
Estrela, R., & Valio, A. 2017, arXiv:1708.05400
D. E. 2010, Science, 329, 187
c(cid:13) 2002 RAS, MNRAS 000, 1–28
Effects of impacts on the TRAPPIST-1 planets
27
Levison, H. F., Morbidelli, A., Tsiganis, K., Nesvorn´y, D.,
Quarles, B., Quintana, E. V., Lopez, E., Schlieder, J. E.,
& Gomes, R. 2011, AJ, 142, 152
& Barclay, T. 2017, ApJ, 842, L5
Lidov, M. L. 1962, Planet. Space Sci. , 9, 719
Lingam, M., & Loeb, A. 2017, Proceedings of the National
Ranjan, S., Wordsworth, R., & Sasselov, D. D. 2017, ApJ,
843, 110
Academy of Science, 114, 6689
Raymond, S. N., Quinn, T., & Lunine, J. I. 2004, Icarus,
Lithwick, Y., & Naoz, S. 2011, ApJ, 742, 94
Liu, B., Munoz, D. J., & Lai, D. 2015, MNRAS, 447, 747
Luger, R., & Barnes, R. 2015, Astrobiology, 15, 119
Luger, R., Sestovic, M., Kruse, E., et al. 2017, Nature As-
tronomy, 1, 0129
MacGregor, M. A., Weinberger, A. J., Wilner, D. J., Kowal-
168, 1
Read, M. J., & Wyatt, M. C. 2016, MNRAS, 457, 465
Rein, H., & Liu, S.-F. 2012, A & A, 537, A128
Rein, H., & Spiegel, D. S. 2015, MNRAS, 446, 1424
Rein, H., & Tamayo, D. 2015, MNRAS, 452, 376
Ribas, I., Bolmont, E., Selsis, F., et al. 2016, A & A, 596,
ski, A. F., & Cranmer, S. R. 2018, ApJ, 855, L2
A111
Marboeuf, U., Bonsor, A., & Augereau, J.-C. 2016,
Planet. Space Sci. , 133, 47
Marino, S., Matr`a, L., Stark, C., et al. 2016, MNRAS, 460,
2933
Rickman, H., Fouchard, M., Froeschl´e, C., & Valsecchi,
G. B. 2008, Celestial Mechanics and Dynamical Astron-
omy, 102, 111
Rickman, H. 2014, Meteoritics and Planetary Science, 49,
Marino, S., Wyatt, M. C., Pani´c, O., et al. 2017, MNRAS,
8
465, 2595
Marino, S., Bonsor, A., Wyatt, M. C., & Kral, Q. 2018,
MNRAS,
Rodler, F., & L´opez-Morales, M. 2014, ApJ, 781, 54
Roettenbacher, R. M., & Kane,
S. R.
2017,
arXiv:1711.02676
Matese, J. J., & Whitman, P. G. 1992, Celestial Mechanics
Rotundi, A., Sierks, H., Della Corte, V., et al. 2015, Sci-
and Dynamical Astronomy, 54, 13
ence, 347, aaa3905
Matr`a, L., Wilner, D. J., Oberg, K. I., et al. 2017, subm.
Matr`a, L., MacGregor, M. A., Kalas, P., et al. 2017, ApJ,
842, 9
Melosh, H. J. 1989, Research supported by NASA. New
York, Oxford University Press (Oxford Monographs on
Geology and Geophysics, No. 11), 1989, 253 p., 11,
Montgomery, S. L., & Welsh, B. Y. 2012, PASP , 124, 1042
Morbidelli, A., Brown, M. E., & Levison, H. F. 2003, Earth
Rugheimer, S., Segura, A., Kaltenegger, L., & Sasselov, D.
2015, ApJ, 806, 137
Saal, A. E., Hauri, E. H., Langmuir, C. H., & Perfit, M. R.
2002, Nature, 419, 451
Saladino, R., Crestini, C., Pino, S., Costanzo, G., & Di
Mauro, E. 2012, Physics of Life Reviews, 9, 84
Schlichting, H. E., Sari, R., & Yalinewich, A. 2015, Icarus,
247, 81
Moon and Planets, 92, 1
Seager, S., Bains, W., & Petkowski, J. J. 2016, Astrobiol-
Morley, C. V., Kreidberg, L., Rustamkulov, Z., Robinson,
ogy, 16, 465
T., & Fortney, J. J. 2017, arXiv:1708.04239
Mumma, M. J., & Charnley, S. B. 2011, ARA&A, 49, 471
Murray, C. D., & Dermott, S. F. 1999, Solar system dynam-
ics by C.D. Murray and S.F. McDermott. (Cambridge,
UK: Cambridge University Press), ISBN 0-521-57295-9
(hc.), ISBN 0-521-57297-4 (pbk.).,
Naoz, S. 2016, ARA&A, 54, 441
Nesvold, E. R., Naoz, S., Vican, L., & Farr, W. M. 2016,
Shannon, A., Wu, Y., & Lithwick, Y. 2016, ApJ, 818, 175
Shuvalov, V. 2009, Meteoritics and Planetary Science, 44,
1095
Shuvalov, V., Kuhrt, E., de Niem, D., & Wunnemann, K.
2014, Planet. Space Sci. , 98, 120
Siever, R. 1968, Sedimentology, 11, 5
Sutherland, J. D. 2017, Nature Reviews Chemistry, , 1,
0012
ApJ, 826, 19
Tamayo, D., Rein, H., Petrovich, C., & Murray, N. 2017,
Nesvorn´y, D., Jenniskens, P., Levison, H. F., et al. 2010,
ApJ, 840, L19
ApJ, 713, 816
Teyssandier, J., Naoz, S., Lizarraga, I., & Rasio, F. A. 2013,
Nesvorn´y, D., & Morbidelli, A. 2012, AJ, 144, 117
Nesvorn´y, D., Roig, F., & Bottke, W. F. 2017, AJ, 153, 103
O'Keefe, J. D., & Ahrens, T. J. 1977, Science, 198, 1249
O'Malley-James, J. T., & Kaltenegger, L. 2017, MNRAS,
469, L26
Ormel, C. W., Ida, S., & Tanaka, H. 2012, ApJ, 758, 80
Ormel, C. W., Liu, B., & Schoonenberg, D. 2017, A & A,
604, A1
Patel, R. I., Metchev, S. A., Heinze, A., & Trollo, J. 2017,
AJ, 153, 54
ApJ, 779, 166
Th´ebault, P., & Augereau, J.-C. 2007, A & A, 472, 169
Theissen, C. A., & West, A. A. 2014, ApJ, 794, 146
Tremaine, S. 1993, Planets Around Pulsars, 36, 335
Turbet, M., Leconte, J., Selsis, F., et al. 2016, A & A, 596,
A112
Unterborn, C. T., Desch, S. J., Hinkel, N. R., & Lorenzo,
A. 2018, Nature Astronomy, 2, 297
Veras, D., & Evans, N. W. 2013, MNRAS, 430, 403
Vitense, C., Krivov, A. V., & Lohne, T. 2010, A & A, 520,
Patel, B. H., Percivalle, C., Ritson, D. J., Duffy, C. D., &
A32
Sutherland, J. D. 2015, Nature Chemistry, 7, 301
Petrovich, C., & Munoz, D. J. 2017, ApJ, 834, 116
Plavchan, P., Werner, M. W., Chen, C. H., et al. 2009, ApJ,
Vitense, C., Krivov, A. V., Kobayashi, H., & Lohne, T.
2012, A & A, 540, A30
Walker, J. C. G., Hays, P. B., & Kasting, J. F. 1981, J. Geo-
698, 1068
Poch, O., Frey, J., Roditi, I., et al. 2017, Astrobiology, 17,
231
Pope, K. O., Baines, K. H., Ocampo, A. C., & Ivanov, B. A.
1997, J. Geophys. Res. , 102, 21645
phys. Res. , 86, 9776
Weidenschilling, S. J. 1977, Ap&SS , 51, 153
Wisdom, J. 1980, AJ, 85, 1122
Wolf, E. T. 2017, ApJ, 839, L1
Wyatt, M. C. 2008, ARA&A, 46, 339
c(cid:13) 2002 RAS, MNRAS 000, 1–28
28 Q. Kral et al.
Wyatt, M. C., Bonsor, A., Jackson, A. P., Marino, S., &
Shannon, A. 2017, MNRAS, 464, 3385
c(cid:13) 2002 RAS, MNRAS 000, 1–28
|
1111.5621 | 1 | 1111 | 2011-11-23T21:00:56 | The Flat Transmission Spectrum of the Super-Earth GJ1214b from Wide Field Camera 3 on the Hubble Space Telescope | [
"astro-ph.EP",
"astro-ph.IM"
] | Capitalizing on the observational advantage offered by its tiny M dwarf host, we present HST/WFC3 grism measurements of the transmission spectrum of the super-Earth exoplanet GJ1214b. These are the first published WFC3 observations of a transiting exoplanet atmosphere. After correcting for a ramp-like instrumental systematic, we achieve nearly photon-limited precision in these observations, finding the transmission spectrum of GJ1214b to be flat between 1.1 and 1.7 microns. Inconsistent with a cloud-free solar composition atmosphere at 8.2 sigma, the measured achromatic transit depth most likely implies a large mean molecular weight for GJ1214b's outer envelope. A dense atmosphere rules out bulk compositions for GJ1214b that explain its large radius by the presence of a very low density gas layer surrounding the planet. High-altitude clouds can alternatively explain the flat transmission spectrum, but they would need to be optically thick up to 10 mbar or consist of particles with a range of sizes approaching 1 micron in diameter. | astro-ph.EP | astro-ph | Draft version December 20, 2013
Preprint typeset using LATEX style emulateap j v. 11/10/09
1
1
0
2
v
o
N
3
2
]
P
E
.
h
p
-
o
r
t
s
a
[
1
v
1
2
6
5
.
1
1
1
1
:
v
i
X
r
a
THE FLAT TRANSMISSION SPECTRUM OF THE SUPER-EARTH GJ1214B FROM WIDE FIELD CAMERA 3
ON THE HUBBLE SPACE TELESCOPE
Zachory K. Berta1 , David Charbonneau1 , Jean-Michel D´esert1 , Eliza Miller-Ricci Kempton2,8 ,
Peter R. McCullough3,4 , Christopher J. Burke5 , Jonathan J. Fortney2 , Jonathan Irwin1 , Philip Nutzman2 ,
Derek Homeier6,7
Draft version December 20, 2013
ABSTRACT
Capitalizing on the observational advantage offered by its tiny M dwarf host, we present HST/WFC3
grism measurements of the transmission spectrum of the super-Earth exoplanet GJ1214b. These are
the first published WFC3 observations of a transiting exoplanet atmosphere. After correcting for a
ramp-like instrumental systematic, we achieve nearly photon-limited precision in these observations,
finding the transmission spectrum of GJ1214b to be flat between 1.1 and 1.7 µm. Inconsistent with a
cloud-free solar composition atmosphere at 8.2σ , the measured achromatic transit depth most likely
implies a large mean molecular weight for GJ1214b’s outer envelope. A dense atmosphere rules out
bulk compositions for GJ1214b that explain its large radius by the presence of a very low density
gas layer surrounding the planet. High-altitude clouds can alternatively explain the flat transmission
spectrum, but they would need to be optically thick up to 10 mbar or consist of particles with a range
of sizes approaching 1 µm in diameter.
Subject headings: planetary systems: individual (GJ 1214b) — eclipses — techniques: spectroscopic
1.
INTRODUCTION
With a radius of 2.7 R⊕ and a mass of 6.5 M⊕ , the
transiting planet GJ1214b (Charbonneau et al. 2009) is
a member of the growing population of exoplanets whose
masses and radii are known to be between those of Earth
and Neptune (see L´eger et al. 2009; Batalha et al. 2011;
Lissauer et al. 2011; Winn et al. 2011). Among these
exoplanets, most of which exhibit such shallow transits
that they require ultra-precise space-based photometry
simply to detect the existence of their transits, GJ1214b
is unique. The diminutive 0.21 R(cid:12) radius of its M dwarf
stellar host means GJ1214b exhibits a large 1.4% tran-
sit depth, and the system’s proximity (13 pc) means the
star is bright enough in the near infrared (H = 9.1) that
follow-up observations to study the planet’s atmosphere
are currently feasible. In this work, we exploit this ob-
servational advantage and present new measurements of
the planet’s atmosphere, which bear upon models for its
interior composition and structure.
According to theoretical studies (Seager et al. 2007;
Rogers & Seager 2010; Nettelmann et al. 2011),
GJ1214b’s 1.9 g cm−3 bulk density is high enough to
[email protected]
1 Harvard-Smithsonian Center for Astrophysics, 60 Garden
St., Cambridge, MA 02138, USA
2 Department of Astronomy and Astrophysics, University of
California, Santa Cruz, CA 95064, USA
3 Space Telescope Science Institute, 3700 San Martin Drive,
Baltimore, MD 21218, USA
4 Smithsonian Astrophysical Observatory, 60 Garden St.,
Cambridge, MA 02138, USA
5 SETI Institute/NASA Ames Research Center, M/S 244-30,
Moffett Field, CA 94035, USA
6 Centre de Recherche Astrophysique de Lyon, UMR 5574,
CNRS, Universit´e de Lyon, ´Ecole Normale Sup´erieure de Lyon,
46 All´ee d’Italie, F-69364 Lyon Cedex 07, France
7 Institut
fur Astrophysik, George-August-Universitat,
Friedrich-Hund-Platz 1, 37077 Gottingen, Germany
8 Sagan Fellow
require a larger ice or rock core fraction than the solar
system ice giants but far too low to be explained with an
entirely Earth-like composition. Rogers & Seager (2010)
have proposed three general scenarios consistent with
GJ1214b’s large radius, where the planet could (i) have
accreted and maintained a nebular H2 /He envelope atop
an ice and rock core, (ii) consist of a rocky planet with
an H2 -rich envelope that formed by recent outgassing, or
(iii) contain a large fraction of water in its interior sur-
rounded by a dense H2 -depleted, H2O-rich atmosphere.
Detailed thermal evolution calculations by Nettelmann
et al. (2011) disfavor this last model on the basis that
it would require unreasonably large bulk water-to-rock
ratios, arguing for at least a partial H2/He envelope, al-
beit one that might be heavily enriched in H2O relative
to the primordial nebula.
By measuring GJ1214b’s transmission spectrum, we
can empirically constrain the mean molecular weight of
the planet’s atmosphere, thus distinguishing among these
possibilities. When the planet passes in front of its
host M dwarf, a small fraction of the star’s light passes
through the upper layers of the planet’s atmosphere be-
fore reaching us; the planet’s transmission spectrum is
then manifested in variations of the transit depth as a
function of wavelength. The amplitude of the transit
depth variations ∆D(λ) in the transmission spectrum
scale as nH × 2H Rp/R2
(cid:63) , where nH is set by the opacities
involved and can be 1-10 for strong absorption features,
H is the atmospheric scale height, Rp is the planetary
radius, and R(cid:63) is the stellar radius (e.g. Seager & Sas-
selov 2000; Brown 2001; Hubbard et al. 2001). Because
the scale height H is inversely proportional to the mean
molecular weight µ of the atmosphere, the amplitude
of features seen in the planet’s transmission spectrum
places strong constraints on the possible values of µ and,
in particular, the hydrogen/helium content of the atmo-
sphere (Miller-Ricci et al. 2009).
2
Berta et al.
Indeed, detailed radiative transfer simulations of
GJ1214b’s atmosphere (Miller-Ricci & Fortney 2010)
show that a solar composition, H2 -dominated atmo-
sphere (µ = 2.4) would show depth variations of roughly
0.1% between 0.6 and 10 µm, while the features in an
H2O-dominated atmosphere (µ = 18) would be an order
of magnitude smaller. While the latter of these is likely
too small to detect directly with current instruments, the
former is at a level that has regularly been measured with
the Hubble Space Telescope (HST) in the transmission
spectra of hot Jupiters (e.g. Charbonneau et al. 2002;
Pont et al. 2008; Sing et al. 2011).
Spectroscopic observations by Bean et al. (2010) with
the Very Large Telescope found the transmission spec-
trum of GJ1214b to be featureless between 0.78-1.0 µm,
down to an amplitude that would rule out cloud-free H2 -
rich atmospheric models. Broadband Spizer Space Tele-
scope photometric transit measurements at 3.6 and 4.5
µm by D´esert et al. (2011a) showed a flat spectrum con-
sistent with Bean et al. (2010), as did high-resolution
spectroscopy with NIRSPEC between 2.0 and 2.4 µm by
Crossfield et al. (2011). Intriguingly, the transit depth
in K -band (2.2 µm) was measured from CFHT by Croll
et al. (2011) to be 0.1% deeper than at other wavelengths,
which would imply a H2 -rich atmosphere, in apparent
contradiction to the other studies.
These seemingly incongruous observations could po-
tentially be brought into agreement if GJ1214b’s at-
mosphere were H2 -rich but significantly depleted in
CH4 (Crossfield et al. 2011; Miller-Ricci Kempton et al.
2011).
In such a scenario, the molecular features that
remain (predominantly H2O) would fit the CFHT mea-
surement, but be unseen by the NIRSPEC and Spitzer
observations. Explaining the flat VLT spectrum in this
context would then require a broadband haze to smooth
the spectrum at shorter wavelengths (see Miller-Ricci
Kempton et al. 2011). New observations by Bean et al.
(2011) covering 0.6-0.85 µm and 2.0-2.3 µm were again
consistent with a flat spectrum, but they still could not
directly speak to this possibility of a methane-depleted,
H2 -rich atmosphere with optically scattering hazes.
Here, we present a new transmission spectrum of
GJ1214b spanning 1.1 to 1.7 µm, using the infrared slit-
less spectroscopy mode on the newly installed Wide Field
Camera 3 (WFC3) aboard the Hubble Space Telescope
(HST). Our WFC3 observations directly probe the pre-
dicted strong 1.15 and 1.4 µm water absorption features
in GJ1214b’s atmosphere (Miller-Ricci & Fortney 2010)
and provide a stringent constraint on the H2 content of
GJ1214b’s atmosphere that is robust to non-equilibrium
methane abundances and hence a definite test of the
CH4 -depleted hypothesis. The features probed by WFC3
are the same features that define the J and H band win-
dows in the telluric spectrum, and cannot be observed
from the ground.
Because this is the first published analysis of WFC3 ob-
servations of a transiting exoplanet, we include a detailed
discussion of the performance of WFC3 in this observa-
tional regime and the systematic effects that are inherent
to the instrument. Recent work on WFC3’s predecessor
NICMOS (Burke et al. 2010; Gibson et al. 2011a) has
highlighted the importance of characterizing instrumen-
tal systematics when interpreting exoplanet results from
HST observations.
Figure 1. A 512x100 pixel cutout of a typical WFC3 G141 grism
exposure of the star GJ1214. The 0th and 1st order spectra are
labeled, and the start of the 2nd order spectrum is visible on the
right. The location of the star in the direct images (not shown
here) is marked with a circle.
This paper is organized as follows: we describe our ob-
servations in §2, our method for extracting spectropho-
tometric light curves from them in §3, and our analy-
sis of these light curves in §4. We present the resulting
transmission spectrum and discuss its implications for
GJ1214b’s composition in §5, and conclude in §6.
2. OBSERVATIONS
We observed three transits of GJ1214b on UT 2010
October 8, 2011 March 28, and 2011 July 23 with the
G141 grism on WFC3’s infrared channel (HST Proposal
#GO-12251, P.I. = Z. Berta), obtaining simultaneous
multiwavelength spectrophotometry of each transit be-
tween 1.1 and 1.7 µm. WFC3’s IR channel consists of a
1024 × 1024 pixel Teledyne HgCdTe detector with a 1.7
µm cutoff that can be paired with any of 15 filters or 2
low-resolution grisms (Dressel et al. 2010). Each expo-
sure is compiled from multiple non-destructive readouts
and can consist of either the full array or a concentric,
smaller subarray.
Each visit consisted of four 96 minute long HST orbits,
each containing 45 minute gaps due to Earth occulta-
tions. Instrumental overheads between the occultations
are dominated by serial downloads of the WFC3 image
buffer, during which all science images are transferred to
the telescope’s solid state recorder. This buffer can hold
only two 16-readout, full-frame IR exposures before re-
quiring a download, which takes 6 minutes. Exposures
cannot be started nor stopped during a buffer download,
so parallel buffer downloads are impossible for short ex-
posures.
Sub ject to these constraints and the possible readout
sequences, we maximized the number of photons detected
per orbit while avoiding saturation by gathering expo-
sures using the 512 × 512 subarray with the RAPID
NSAMP=7 readout sequence, for an effective integration
time of 5.971 seconds per exposure. With this setting,
four 12-exposure batches, separated by buffer downloads,
were gathered per orbit resulting in an integration effi-
ciency of 10%. Although the brightest pixel in the 1st or-
der spectrum reaches 78% of saturation during this expo-
sure time, the WFC3’s multiple non-destructive readouts
enable the flux within each pixel to be estimated before
the onset of significant near-saturation nonlinearities.
A sub-region of a typical G141 grism image of GJ1214
is shown in Fig. 1. The 512 × 512 subarray allows both
the 0th and 1st order spectra to be recorded, and the
1st order spectrum to fall entirely within a single ampli-
fier quadrant of the detector. The 1st order spectrum
spans 150 pixels with a dispersion in the x-direction of
4.65 nm/pixel and a spatial full-width half maximum in
the y -direction of 1.7 pixels (0.2”). The 0th order spec-
trum is slightly dispersed by the grism’s prism but is
nearly a point source. Other stars are present in the
1st order0th orderdirectimageWFC3 Observations of GJ1214b
3
that calwf3 does not apply flat-field corrections when
calibrating grism images; proper wavelength-dependent
flat-fielding for slitless spectroscopy requires wavelength-
calibrating individual sources and calwf3 does not per-
form this task.
3.1. Interpolating over Cosmic Rays
calwf3 identifies cosmic rays that appear partway
through an exposure by looking for deviations from a
linear accumulation of charge among the non-destructive
readouts, but it can not identify cosmic rays that appear
between the zeroth and first readout. We supplement
calwf3’s cosmic ray identifications by also flagging any
pixel in an individual exposure that is > 6σ above the
median of that pixel’s value in all other exposures as a
cosmic ray. Through all three visits (576 exposures), a
total of 88 cosmic rays were identified within the extrac-
tion box for the 1st order spectra.
For each exposure, we spatially interpolate over cos-
mic rays. Near the 1st order spectrum, the pixel-to-pixel
gradient of the point spread function (PSF) is typically
much shallower along the dispersion direction than per-
pendicular to it, so we use only horizontally adjacent
pixels when interpolating to avoid errors in modeling the
sharp cross-dispersion falloff.
3.2. Identifying Continuously Bad Pixels
We also mask any pixels that are identified as “bad
detector pixels” (DQ=4), “unstable response” (DQ=32),
“bad or uncertain flat value” (DQ=512). We found that
only these DQ flags affected the photometry in a pixel
by more than 1σ . Other flags may have influenced the
pixel photometry, but did so below the level of the pho-
ton noise. In the second visit, we also identified one col-
umn of the detector (x = 625 in physical pixels9 ) whose
light curve exhibited a dramatically different systematic
variation than did light curves from any of the other
columns. This column was coincident with an unusually
low-sensitivity feature in the flat-field, and we hypothe-
size that the flat-field is more uncertain in this column
than in neighboring columns. We masked all pixels in
that column as bad.
We opt not to interpolate over these continuously bad
pixels. Because they remained flagged throughout the
duration of each visit, we simply give these pixels zero
weight when extracting 1D spectra from the images. This
allows us to keep track of the actual number of photons
recorded in each exposure so we can better assess our
predicted photometric uncertainties.
3.3. Background Estimation
In addition to the target, WFC3 also detects light from
the diffuse sky background, which comes predominantly
from zodiacal light and Earth-shine, and must be sub-
tracted. We draw conservative masks around all sources
that are visible in each visit’s median image, including
GJ1214 and its electronics cross-talk artifact (see Viana
& Baggett 2010). We exclude these pixels, as well as all
9 For ease of comparison with future WFC3 analyses, through-
out this paper we quote all pixel positions in physical units as
interpreted by SAOImage DS9, where the bottom left pixel of a
full-frame array would be (x, y) = (1,1).
Figure 2. The mean out-of-transit extracted spectrum of GJ1214
(black line) from all three HST visits, shown before (top) and af-
ter (bottom) flux calibration.
Individual extracted spectra from
each visit are shown with their 1σ uncertainties (color error bars).
For comparison, the integrated flux from the PHOENIX model at-
mosphere used to calculate the stellar limb darkening (see §4.2) is
shown (gray lines) offset for clarity and binned to the WFC3 pixel
scale (gray circles).
subarray’s 68(cid:48) × 61(cid:48) field of view, but are too faint to
provide useful diagnostics of systematic trends that may
exist in the data. For wavelength calibration, we gath-
ered direct images in the F130N narrow-band filter; the
direct images’ position relative to the grism images is
also shown in Fig. 1.
To avoid systematics from the detector flat-fields that
have a quoted precision no better than 0.5% (Pirzkal
et al. 2011), the telescope was not dithered during any
of the observations. We note that a technique called
“spatial scanning” has been proposed to decrease the
overheads for bright targets with WFC3, where the tele-
scope nods during an exposure to smear the light along
the cross-dispersion direction, thus increasing the time
to saturation (McCullough & MacKenty 2011). We did
not use this mode of observation as it was not yet tested
at the time our program was initiated.
3. DATA REDUCTION
The Python/PyRAF software package aXe was devel-
oped to extract spectra from slitless grism observations
with WFC3 and other Hubble instruments (Kummel
et al. 2009), but it is optimized for extracting large num-
bers of spectra from full frame dithered grism images.
To produce relative spectrophotometric measurements of
our single bright source, we opted to create our own ex-
traction pipeline that prioritizes precision in the time
domain. We outline the extraction procedure below.
Through the
extraction, we use
calibrated 2-
dimensional images, the “flt” outputs from WFC3’s
calwf3 pipeline. For each exposure, calwf3 performs the
following steps: flag detector pixels with the appropriate
data quality (DQ) warnings, estimate and remove bias
drifts using the reference pixels, subtract dark current,
determine count rates and identify cosmic rays by fitting
a slope to the non-destructive reads, correct for photo-
metric non-linearity (properly accounting for the signal
accumulation before the initial “zeroth” read), and apply
gain calibration. The resulting images are measured in
e− s−1 and contain per pixel uncertainty estimates based
on a detector model (Kim Quijano et al. 2009). We note
4
Berta et al.
a simple mean of the unmasked pixels; the results were
unchanged.
3.4. Inter-pixel Capacitance
The normal calibration pipeline does not correct for
the inter-pixel capacitance (IPC) effect, which effectively
couples the flux recorded in adjacent pixels at about
the 1% level (McCullough 2008). We correct this ef-
fect with a linear deconvolution algorithm (McCullough
2008; Hilbert & McCullough 2011), although we find it
makes little difference to the final results.
3.5. Extracting the Zeroth Order Image
The 0th order image can act as a diagnostic for tracking
changes in the telescope pointing and in the shape of the
instrumental PSF. We select a 10 × 10 pixel box around
the 0th order image and fit a 2D Gaussian to it with the
x position, y position, size in the x direction, size in the y
direction, and total flux allowed to vary (5 parameters).
Time series of the 0th order x and y positions, sizes
in both directions, and total flux are shown for all three
visits in Fig. 3 (panels c-g). Thanks to the dispersion by
the grism’s prism, the Gaussian is typically 20% wider in
the x direction than in the y direction. Even though the
throughput of the 0th order image is a factor of 60 lower
than the 1st order spectrum, the transit of GJ1214b is
readily apparent in the 0th -order flux time series.
3.6. Extracting the First Order Spectrum
To extract the first order spectra, we first determine
the position of GJ1214 in the direct image, which serves
as a reference position for defining the trace and wave-
length calibration of the 1st order spectrum. We adopt
the mean position GJ1214 in all of the direct images as
the reference position, which we measure using the same
method as in extracting the 0th order image in §3.5. The
measured (x, y) reference positions for the first, second,
and third visits are (498.0, 527.5), (498.6, 531.1), and
(498.9, 527.1) in physical pixels.
Once the reference pixel for a visit is known, we use
the coefficients stored in the WFC3/G141 aXe configu-
ration file10 (Kuntschner et al. 2009), to determine the
geometry of the 1st order trace and cut out a 30 pixel
tall extraction box centered on the trace. Within this
extraction box, we use the wavelength calibration coeffi-
cients to determine the average wavelength of light that
will be illuminating each pixel. We treat all pixels in the
same column as having the same effective wavelength;
given the spectrum’s 0.5◦ tilt from to the x axis, errors
introduced by this simplification are negligible.
Kuntschner et al. (2008) used flat-fields taken through
all narrow-band filters available on WFC3/IR to con-
struct a flat-field “cube” where each pixel contains 4
polynomial coefficients that describe its sensitivity as a
function of wavelength11 . We use this flat-field cube to
construct a color-dependent flat based on our estimate of
the effective wavelength illuminating each pixel, and di-
vide each exposure by it. WFC3 wavelength calibration
and flat-fielding is described in detail in the aXe manual
(Kummel et al. 2010).
10 The aXe configuration file WFC3.IR.G141.V2.0.conf is avail-
able through http://www.stsci.edu/hst/wfc3/
11 WFC3.IR.G141.flat.2.fits, through the same URL
Figure 3. Extracted properties of the 0th and 1st order spectra as
a function of time, including (a) the summed 1st order photometric
light curves; (b) the estimated sky background level; (c-g) the total
flux, x and y position (measured relative to the reference pixel),
and Gaussian widths in the x and y directions of the 0th order
image; and (h-j) the y offset, cross-dispersion width, and slope of
the 1st order spectrum. All three visits are shown and are denoted
by the color of the symbols. The 45 minute gaps in each time series
are due to Earth occultations, the 6 minute gaps are due to the
WFC3 buffer downloads.
pixels that have any DQ warning flagged. Then, to es-
timate the sky background in each exposure, we scale a
master WFC3 grism sky image (Kummel et al. 2011) to
match the remaining 70-80% of the pixels in each expo-
sure and subtract it. We find typical background levels
of 1 − 3 e− s−1 pixel−1 , that vary smoothly within or-
bits and throughout visits as shown in Fig. 3 (panel b).
As a test, we also estimated the background level from
WFC3 Observations of GJ1214b
5
To calculate 1D spectra from the flat-fielded images,
we sum all the unmasked pixels within the extraction
box over the y -axis. To estimate the uncertainty in each
spectral channel, we first construct a per-pixel uncer-
tainty model that includes photon noise from the source
and sky as well as 22 e− of read noise, and sum these
uncertainties, in quadrature, over the y -axis. We do not
use the calwf3-estimated uncertainties; they include a
term propagated from the uncertainty in the nonlinearity
correction that, while appropriate for absolute photome-
try, would not be appropriate for relative photometry. In
each exposure, there are typically 1.2 × 105 e− per single-
pixel spectral channel and a total of 1.5 × 107 e− in the
entire spectrum. Fig. 3 (panel a) shows the extracted
spectra summed over all wavelengths as a function of
time, the “white” light curve.
For diagnostics’ sake, we also measure the geometri-
cal properties of the 1st order spectra in each exposure.
We fit 1D Gaussians the cross-dispersion profile in each
column of the spectrum and take the median Gaussian
width among all the columns as a measurement of the
PSF’s width. We fit a line to the location of the Gaus-
sian peaks in all the columns, taking the intercept and
the slope of that line as an estimate of the y -offset and
tilt of the spectrum on the detector. Time series of these
parameters are shown in Fig. 3 (panels h-j).
3.7. Flux Calibration
For the sake of display purposes only (see Fig. 2), we
flux calibrate each visit’s median, extracted, 1D spec-
trum. Here we have interpolated over all bad pixels
within each visit (contrary to the discussion in the §3.2),
and plotted the weighted mean over all three visits. The
calibration uncertainty for the G141 sensitivity curve
(Kuntschner et al. 2011) is quoted to be 1%.
3.8. Times of Observations
For each exposure, we extract the EXPSTART keyword
from the science header, which is the Modified Julian
Date at the start of the exposure. We correct this to
the mid-exposure time using the EXPTIME keyword, and
convert it to the Barycentric Julian Date in the Barycen-
tric Dynamical Time standard using the code provided
by Eastman et al. (2010).
4. ANALYSIS
In this section we describe our method for estimating
parameter uncertainties (§4.1) and our strategy for mod-
eling GJ1214’s stellar limb darkening (§4.2). Then, af-
ter identifying the dominant systematics in WFC3 light
curves (§4.3) and describing a method to correct them
(§4.4), we present our fits to the light curves, both
summed over wavelength (§4.5) and spectroscopically re-
solved (§4.6). We also present a fruitless search for tran-
siting satellite companions to GJ1214b (§4.7).
4.1. Estimating Parameter Distributions
Throughout our analysis, we fit different WFC3 light
curves with models that have different sets of parame-
ters, and draw conclusions from the inferred probability
distributions of those parameters; this section describes
our method for characterizing the posterior probability
distribution for a set of parameters within a given model.
=
(1)
We use a Markov Chain Monte Carlo (MCMC) method
with the Metropolis-Hastings algorithm to explore the
posterior probability density function (PDF) of the
model parameters. This Bayesian technique allows us
to sample from (and thus infer the shape of ) the prob-
ability distribution of a model’s parameters given both
our data and our prior knowledge about the parameters
(for reviews, see Ford 2005; Gregory 2005; Hogg et al.
2010). Briefly, the algorithm starts a chain with an ini-
tial set of parameters (Mj=0 ) and generates a trial set of
parameters (M(cid:48)
j+1 ) by perturbing the previous set. The
ratio of posterior probability between the two parameter
sets, given the data D, is then calculated as
j+1 D)
P (DM(cid:48)
P (M(cid:48)
× P (M(cid:48)
j+1 )
j+1 )
P (DMj )
P (Mj D)
P (Mj )
where the first term (the “likelihood”) accounts for the
information that our data provide about the parameters
and the second term (the “prior”) specifies our externally
conceived knowledge about the parameters. If a random
number drawn from a uniform distribution between 0
and 1 is less than this probability ratio, then Mj+1 is set
to M(cid:48)
j+1 ; if not, then Mj+1 reverts to Mj . The process
is iterated until j is large, and the resulting chain of
parameter sets is a fair sample from the posterior PDF
and can be used to estimate confidence intervals for each
parameter.
To calculate the likelihood term in Eq. 1, we assume
that each of the N flux values di is drawn from a uncorre-
lated Gaussian distribution centered on the model value
mi with a standard deviation of sσi , where σi is the the-
oretical uncertainty for the flux measurement based on
the detector model and photon statistics and s is a pho-
tometric uncertainty rescaling parameter. Calculation of
the ratio in Eq. 1 is best done in logarithmic space for
numerical stability, so we write the likelihood as
ln P (DM) = −N ln s − 1
2s2 χ2 + constant
(cid:18) di − mi
(cid:19)2
N(cid:88)
σi
i=1
where
χ2 =
(2)
(3)
and we have only explicitly displayed terms that depend
on the model parameters.
Including s as a model pa-
rameter is akin to rescaling the uncertainties by exter-
nally modifying σi to achieve a reduced χ2 of unity, but
enables the MCMC to fit for and marginalize over this
rescaling automatically. Unless otherwise stated for spe-
cific parameters, we use non-informative (uniform) priors
for the second term in Eq. 1. We use a Jeffreys prior on s
(uniform in ln s) which is the least informative, although
the results are practically indistinguishable from prior
uniform in s.
When generating each new trial parameter set M(cid:48)
j+1 ,
we follow Dunkley et al. (2005) and perturb every pa-
rameter at once, drawing the parameter jumps from a
multivariate Gaussian with a covariance matrix that ap-
proximates that of the parameter distribution. Doing so
allows the MCMC to move easily along the dominant lin-
ear correlations in parameter PDF, and greatly increases
the efficiency of the algorithm. While this procedure may
6
Berta et al.
Table 1
White Light Curves from WFC3/G141
Timea
(BJDTDB )
2455478.439980
2455478.440270
2455478.440559
2455478.440848
2455478.441138
Relative Fluxb Uncertainty
0.99381
0.99713
0.99787
0.99958
0.99989
0.00031
0.00031
0.00031
0.00032
0.00032
Sky
(e− /s)
1.9546
1.9938
1.9718
1.8808
1.9379
0th -Xc
(pix)
0th -Yc
(pix)
0th -Ad
(pix)
0th -Bd
(pix)
1st -Yc
(pix)
-187.830
-187.839
-187.846
-187.827
-187.844
...
-0.474
-0.481
-0.484
-0.490
-0.490
0.790
0.792
0.789
0.792
0.785
0.613
0.616
0.615
0.614
0.612
-0.075
-0.082
-0.076
-0.087
-0.085
1st -Bd
(pix)
0.7484
0.7490
0.7486
0.7476
0.7492
1st -Slopee Visit
(pix/pix)
0.00921
0.00925
0.00914
0.00917
0.00921
1
1
1
1
1
Note. — This table is published in its entirety in the electronic edition of the Astrophysical Journal. A portion is
shown here for guidance concerning its form and content.
a Mid-exposure time.
b Normalized to the median flux level of the out-of-transit observations in each visit.
c Position measured relative to the Gaussian center of each visit’s direct image.
d Gaussian width of the 0th or 1st order spectra in the horizontal (A) or vertical (B) direction.
e Slope of the 1st order spectrum.
seem circular (if we knew the covariance matrix of the pa-
rameter distribution, why would we need to perform the
MCMC?), the covariance matrix we use to generate trial
parameters could be a very rough approximation to the
true shape of the parameter PDF but still dramatically
decrease the computation time necessary for the MCMC.
To obtain an initial guess for parameters (Mj=0 ),
we use the MPFIT implementation (Markwardt 2009)
of the Levenberg-Marquardt (LM) method to maximize
ln P (MD). This would be identical to minimizing χ2
in the case of flat priors, but it can also include con-
straints from more informative priors. The LM fit also
provides an estimate of the covariance matrix of the pa-
rameters, which is a linearization of the probability space
near the best-fit. We use this covariance matrix esti-
mate for generating trial parameters in the MCMC, and
with it, achieve parameter acceptance rates of 10-40%
throughout the following sections. As expected, when
fitting models with flat priors and linear or nearly-linear
parameters (where the PDF should well-described by
a multivariate Gaussian), the LM covariance matrix is
identical to that ultimately obtained from the MCMC
(see Sivia 1996, for further discussion).
MCMC chains are run until they contain 1.25 × 105
points. The first 1/5 of the points are ignored as “burn-
in”, leaving 1×105 for parameter estimation. Correlation
lengths for the parameters in the MCMC chains are indi-
cated throughout the text; they are typically of order 10
points. A chain with such a correlation length effectively
contains 1 × 105/10 = 1 × 104 independent realizations
of the posterior PDF. We quote confidence intervals that
exclude the upper and lower 16% of the marginalized dis-
tribution for each parameter (i.e. the parameter’s central
68% confidence interval), using all 1×105 points in each
chain.
4.2. Modeling Stel lar Limb Darkening
Accurate modeling of the WFC3 integrated and spec-
troscopic transit light curves requires careful consid-
eration of the stellar limb-darkening (LD) behavior.
GJ1214b’s M4.5V stellar host is so cool that it exhibits
weak absorption features due to molecular H2O. Because
inferences of the planet’s apparent radii from transit light
curves depend strongly on the star’s limb-darkening,
which is clearly influenced by H2O as an opacity source,
inaccurate treatment of limb-darkening could potentially
introduce spurious H2O features into the transmission
spectrum.
If they were sufficiently precise, transit light curves
alone could simultaneously constrain both the star’s mul-
tiwavelength limb-darkening behavior and the planet’s
multiwavelength radii (e.g. Knutson et al. 2007b). For
less precise light curves, it is common practice to fix the
limb-darkening to a theoretically calculated law, even if
this may underestimate the uncertainty in the planetary
parameters (see Burke et al. 2007; Southworth 2008).
Given the quality of our data, we adopt an intermediate
solution where we allow the limb-darkening parameters
to vary in our fits, but with a Gaussian prior centered on
the theoretical values (e.g. Bean et al. 2010).
We model the star GJ1214’s limb-darkening behav-
ior with a spherically symmetric PHOENIX atmosphere
(Hauschildt et al. 1999), assuming stellar parameters of
Teff = 3026K , log g = 5, and [M/H] = 0 (Charbonneau
et al. 2009). As shown in Fig. 2, the integrated flux
from the PHOENIX model is in good qualitative agree-
ment with the low-resolution, calibrated WFC3 stellar
spectrum of GJ1214. From this model, we calculate
photon-weighted average intensity profiles for the inte-
grated spectrum and for each of the individual wave-
length bins, using the WFC3 grism sensitivity curve and
the PHOENIX model to estimate the photon counts.
In
the spherical geometry of the PHOENIX atmospheres
the characterization of the actual limb (defined as µ = 0,
see below) is not straightforward, as the model extends
beyond the photosphere into the optically thin outer at-
mosphere. The result is an approximately exponentially
declining intensity profile from the outermost layers, that
Claret & Hauschildt (2003) found not to be easily repro-
duced by standard limb darkening laws for plane-parallel
atmospheres. These authors suggest the use of “quasi-
spherical” models by ignoring the outer region.
In an
extension of this concept, we set the outer surface of the
star to be where the intensity drops to e−1 of the central
intensity, and measure µ = cos θ (where θ is the emission
angle relative to the line of sight) relative to that outer
radius.
We derive coefficients for a square-root limb-darkening
WFC3 Observations of GJ1214b
7
sures between buffer downloads. To the eye, the ramps
are very repeatable; the flux at the end of all batches
asymptotes to nearly the same level. The amplitude of
the ramp is 0.4% from start to finish for most batches,
except for the first batch of each orbit, where the ramp
is somewhat less pronounced.
These ramps are reminiscent of those seen in high-
cadence Spitzer light curves at 8 and 16 µm (e.g. Dem-
ing et al. 2006; Knutson et al. 2007a; Charbonneau et al.
2008) which Agol et al. (2010) recently proposed may
be due to “charge trapping” within the detector pixels.
In their toy model, charge traps within each pixel be-
come filled throughout an exposure and later release the
trapped charge on a finite timescale, thereby increasing
the pixel’s dark current in subsequent exposures. The
model leads to exponential ramps when observing bright
sources as the excess dark current increases sharply at
first but slows its increase as the population of charge
traps begins to approach steady state. We note this
model also leads to after-images following strong expo-
sures, i.e., persistence.
WFC3 has been known since its initial ground-testing
to exhibit strong persistence behavior (McCullough &
Deustua 2008; Long et al. 2010). Smith et al. (2008a)
have proposed that persistence in 1.7 µm cutoff HgCdTe
detectors like WFC3 is likely related to charge trapping.
Measurements (McCullough & Deustua 2008) indicate
that WFC3’s persistence may be of the right order of
magnitude (on < 1 minute timescales) to supply the
roughly 50 e− s−1 pixel−1 in the brightest pixels that
would be necessary to explain the observed several mil-
limagnitude ramp, although persistence levels and de-
cay timescales can depend in complicated ways on the
strength of previous exposures (see Smith et al. 2008b).
We were aware of this persistence issue before our ob-
servations and made an effort to control its effect on our
light curves. When we planned the timing of the expo-
sures, we attempted to make the illumination history of
each pixel as consistent as possible from batch to batch
and orbit to orbit. In practice, this means we gathered
more direct images than necessary for wavelength cali-
bration to delay some of the grism exposures.
Whether or not the ramps are caused by the charge
trapping mechanism, they are definitely dependent on
the illumination that a pixel receives. To demonstrate
this, we construct light curves for each individual pixel
over the duration of every out-of-transit 12-exposure
batch that follows a buffer download and normalize each
of these pixel light curves to the first exposure in the
batch. Fig. 4 shows the normalized pixel light curves,
grouped by their mean recorded fluence. Because it takes
a finite time to read the subarray (0.8 seconds) and reset
the full array (2.9 seconds), we note that each exposure
actually collects 60% more electrons than indicated by
these nominal, recorded fluences (see Long et al. 2011).
The appearance of the ramp clearly becomes more pro-
nounced for pixels that are more strongly exposed.
Buried beneath the ramp features, the summed light
curve exhibits subtler trends that appear mostly as orbit-
long or visit-long slopes with a peak-to-peak variation of
about 0.05%. These are perhaps caused by slow drifts
in pointing and focus (telescope “breathing”) interacting
with sensitivity variations across the detector that are
Figure 4. Single-pixel light curves within each 12-exposure batch
following a buffer download (gray points), shown for different mean
pixel illuminations. Pixel light curves have been normalized to the
first exposure within each batch, and plotted with small random
horizontal offsets for clarity. Only data from the first HST visit,
which exhibited the smallest pointing drifts (see Fig. 3), are shown.
Error bars show the mean and its standard error for each time point
and each illumination. An exponential ramp begins is present in
pixels with a mean recorded fluence greater than 30,000 and 40,000
e− (50% of the detector full well). Note, the nominal fluences
quoted here do not include charge accumulated during detector
flushing and initial readout (see text).
law for each of these average intensity profiles using least-
squares fitting. In this law, the intensity relative to the
center of the star is given by
= 1 − c(1 − µ) − d(1 − √
(4)
I (µ)
I (1)
µ),
where c and d are the two coefficients of the fit. We
chose a square-root law over the popular quadratic law
because it gave noticeably better approximations to the
PHOENIX intensity profiles, while still having few enough
free parameters that they can be partially inferred from
the data. Indeed, van Hamme (1993) found the square-
root law to be generally preferable to other 2-parameter
limb-darkening laws for late-type stars in the near-IR.
The square-root law matches the theoretical intensity
profile nearly as well as the full nonlinear 4-parameter
law introduced by Claret (2000) for the models we use
here.
4.3. Light Curve Systematics
The summed light curve shown in Fig. 3 (panel a) ex-
hibits non-astrophysical systematic trends. The most ob-
vious of these are the sharply rising but quickly saturat-
ing “ramp”-like features within each batch of 12 expo-
8
Berta et al.
not perfectly corrected by the flat field.
4.4. Correcting for Systematics
Fortunately, these systematics are extremely repeat-
able between orbits within a visit; we harness this fact
when correcting for them. We divide the in-transit
orbit of any photometric timeseries, either the white
light curve or one of the spectroscopically resolved light
curves, by a systematics correction template constructed
from the two good out-of-transit orbits. This template is
simply the weighted average of the fluxes in the out-of-
transit orbits, evaluated at each exposure within an orbit.
It encodes both variations in the effective sensitivity of
the detector within an orbit and the mean out-of-transit
flux level.
When performing the division, we propagate the tem-
of (cid:112)1 + 1/2 = 1.22. This factor, although it may seem
plate uncertainty into the photometric uncertainty for
each exposure, which typically increases it by a factor
like an undesired degradation of the photometric preci-
sion, would inevitably propagate into measurements of
the transit depth whether we performed this correction
or not, since Rp/R(cid:63) is always measured relative to the
out-of-transit flux, which must at some point be inferred
from the data.
Throughout
to this pro-
refer
this work, we
cess of dividing by the out-of-transit orbits as the
divide-oot method. Because each point in the sin-
gle in-transit orbit is equally spaced in time between
the two out-of-transit exposures being used to correct
it, the divide-oot method also naturally removes the
0.05% visit-long slope seen in the raw photometry. As
we show in §4.5, when applied to the white light curves,
the divide-oot treatment produces uncorrelated Gaus-
sian residuals that have a scatter consistent with the pre-
dicted photon uncertainties.
Unlike decorrelation techniques that have often been
used to correct systematics in HST light curves, the
divide-oot method does not require knowing the re-
lationship between measured photometry and the phys-
ical state of the camera.
It does, however, strictly re-
quire the systematics to repeat over multiple orbits. The
divide-oot method would not work if the changes in
the position, shape, and rotational angle of the 1st -order
spectrum were not repeated in the other orbits in a visit
or if the cadence of the illumination were not nearly iden-
tical across orbits. In such cases, the Gaussian process
method proposed by Gibson et al. (2011b) may be a use-
ful alternative, and one that would appropriately account
for the uncertainty involved in the systematics correc-
tion.
4.5. White Light Curve Fits
Although the main scientific result of this paper is de-
rived from the spectroscopic light curves presented in
§4.6, we also analyze the light curve summed over all
wavelengths between 1.1 and 1.7 µm. We use these
white light curves to confirm the general system prop-
erties found in previous studies and quantitatively inves-
tigate the instrumental systematics.
We fit an analytic, limb-darkened transit light curve
model (Mandel & Agol 2002) to the divide-oot-
corrected white light curves. Only the in-transit orbits
were fit; after the divide-oot correction, the two out-
of-transit orbits contain no further information. Also,
because the in-transit orbit’s flux has already been nor-
malized, we fix the out-of-transit flux level to unity
in all the fits. Throughout, we fix the planet’s pe-
riod to P = 1.58040481 days and mid-transit time to
Tc = 2454966.525123 BJDTDB (Bean et al. 2011), the
orbital eccentricity to e = 0, and the stellar mass to
0.157M(cid:12) (Charbonneau et al. 2009).
4.5.1. Combined White Light Curve
First, we combine the three visits into a single light
curve, as shown in Fig. 5, and fit for the following pa-
rameters: the planet-to-star radius ratio (Rp/R(cid:63) ), the
total transit duration between first and fourth contact
(t14 ), the stellar radius (R(cid:63) ), and the two coefficients c
and d of the square-root limb-darkening law12 . Previous
studies have found no significant transit timing variations
for the GJ1214b system (Charbonneau et al. 2009; Sada
et al. 2010; Bean et al. 2010; Carter et al. 2011; D´esert
et al. 2011a; Kundurthy et al. 2011; Berta et al. 2011;
Croll et al. 2011), so we fix the time of mid-transit for
each visit to be that predicted by the linear ephemeris.
As in Burke et al. (2007), we use the parameters t14
and R(cid:63) to ensure quick convergence of the MCMC be-
cause correlations among these parameters are more lin-
ear than for the commonly fit impact parameter (b) and
scaled semi-ma jor axis (a/R(cid:63) ). Because nonlinear trans-
formations between parameter pairs will deform the hy-
pervolume of parameter space, we include a Jacobian
term in the priors in Eq. 1 to ensure uniform priors for
the physical parameters Rp , R(cid:63) , and i (see Burke et al.
2007; Carter et al. 2008, for detailed discussions). For
the combined light curves, the influence of this term is
practically negligible, but we include it for completeness.
In the MCMC chains described in this section, all pa-
rameters have correlation lengths of 6-13 points.
Initially, we perform the fit with limb-darkening coeffi-
cients c and d without any priors from the PHOENIX atmo-
sphere model, enforcing only that 0 < c + d < 1, which
ensures that the star is brighter at its center (µ = 1)
than at its limb (µ = 0).
Interestingly, the quantity
(c/3 + d/5), which sets the integral of I (µ) over the
stellar surface, defines the line along which c and d are
most strongly correlated in the MCMC samples (see also
Irwin et al. 2011). For quadratic limb-darkening, the
commonly quoted 2u1 + u2 combination (Holman et al.
2006) has the same physical meaning. The integral of
I (µ) can be thought of as the increase in the central
transit depth over that for a constant-intensity stellar
disk, so it makes sense that it is well-constrained for
nearly equatorial transiting systems like GJ1214b. Plan-
ets with higher impact parameters do not sample the full
range of 0 < µ < 1 during transit, leading to correspond-
ingly weaker limb-darkening constraints that can be de-
rived from their light curves (see Knutson et al. 2011).
We quote confidence intervals for the linear combination
(c/3 + d/5) and one orthogonal to it in Table 2, along
with rest of the parameters.
Heartened by finding that when they are allowed to
vary freely, our inferred white-light limb-darkening coef-
12 The square-root law is a special case of the 4-parameter law
and straightforward to include in the Mandel & Agol (2002) model.
WFC3 Observations of GJ1214b
9
Figure 5. The white light curve of GJ1214b’s transits before (top panels) and after (midd le panels) removing the instrumental systematics
using the divide-oot (left) and model-ramp (right, with offsets for clarity) methods described in §4.4 and §4.5.4. A transit model that was
fit to the divide-oot-corrected light curve, constrained to the values of a/R(cid:63) and b used by Bean et al. (2010), is shown (gray lines), along
with residuals from this model (bottom panels). In the left panels, the out-of-transit orbits are not shown after the correction has been
applied, because they contain no further information.
ficients agree to 1σ to those derived using the PHOENIX
stellar model, we perform a second fit that includes the
PHOENIX models as informative priors. For this prior,
we say P (M) in Eq. 1 is proportional to a Gaussian
with (c/3 + d/5) = 0.0892 ± 0.018 and (c/5 − d/3) =
−0.431 ± 0.032, which is centered on the PHOENIX model.
To set the 1σ widths of these priors, we start by vary-
ing the effective temperature of the star in the PHOENIX
model by its 130K uncertainty in either direction, and
then double the width of the prior beyond this, to ac-
count for potential systematic uncertainties in the atmo-
sphere model. The results from the fit with these LD
priors are shown in Table 2.
The photometric noise rescaling parameter s is within
10% of unity, implying that the 376 ppm achieved scat-
ter in the combined white light curve can be quite well-
explained from the known sources of uncertainty in the
measurements, predominantly photon noise from the
star. As shown in Fig. 6, for the divide-oot-corrected
light curves, the autocorrelation function (ACF) of the
residuals shows no evidence for time-correlated noise.
Likewise, the scatter in binned divide-oot residuals de-
creases as the square-root of the number of points in a
bin, as expected for uncorrelated Gaussian noise. If there
are uncorrected systematic effects remaining in the data,
they are below the level of the photon noise over the
time-scales of interest here.
4.5.2. Individual White Light Curves
To test for possible differences among our WFC3 vis-
its, we fit each of the three divide-oot-corrected white
light curves individually. In addition to Rp/R(cid:63) , t14 , and
R(cid:63) , we also allow ∆Tc (the deviation of each visit’s mid-
transit time from the linear ephemeris) to vary freely.
We allow c and d to vary, but enforce the same PHOENIX-
derived priors described in §4.5.1.
Table 3 shows the results, which are consistent with
each other and with other observations (Charbonneau
et al. 2009; Bean et al. 2010; Carter et al. 2011; Kun-
Figure 6. For the transit model in Fig. 5 and both types of sys-
tematics treatments, the autocorrelation function of the residuals
(ACF; top) and the scatter in binned residuals as a function of
bin size (bottom). The residuals from the combined light curve
are shown (black points), as well as the individual visits (colorful
√
points). The expectations from uncorrelated Gaussian noise (0 in
the top, ∝ 1/
N in the bottom) are overplotted (dashed lines).
durthy et al. 2011; Berta et al. 2011). The three mea-
sured ∆Tc ’s show no evidence for transit timing varia-
tions. The uncertainties for the parameters t14 , R(cid:63) , and
∆Tc are noticeably largest in the first visit; this is most
likely because the first visit does not directly measure
the timing of either 1st or 4th contact, on which these
parameters strongly depend. Additionally, whereas the
correlation lengths in the MCMC chains for these pa-
rameters in the two visits that do measure 1st/4th con-
tact and for Rp/R(cid:63) , c, and d in all three visits are small
(10-30 points), the correlation lengths for t14 , R(cid:63) , and
∆Tc in the first visit are very large (300-400 points), in-
dicating these weakly constrained parameters are poorly
approximated by the MPFIT-derived covariance matrix.
On account of the large correlation lengths for these pa-
rameters, we ran the MCMC for the first visit with a
factor of 10 more points. In each of the three visits, the
10
Berta et al.
Table 2
Transit Parameters Inferred from the Combined White Light
Curve
Parameter
t14 (days)
R(cid:63) (R(cid:12) )
a/R(cid:63)
i (◦ )
b
Rp /R(cid:63)
c/3 + d/5
c/5 − d/3
b
predicted RMS
achieved RMS
s
No LD Prior
0.03624 ± 0.00013
0.2014+0.0038
−0.0025
15.30+0.19−0.29
89.3 ± 0.4
0.18+0.09−0.11
0.1158+0.0007
−0.0006
0.096 ± 0.008
−0.52+0.22−0.14
337 ppm
373 ppm
1.12 ± 0.07
LD Priora
0.03620 ± 0.00012
0.201+0.004
−0.003
15.31+0.21−0.29
89.3+0.4−0.3
0.19+0.08−0.11
0.1160 ± 0.0005
0.095 ± 0.007
−0.433 ± 0.032
337 ppm
376 ppm
1.12 ± 0.07
a The Gaussian limb-darkening priors of (c/3 + d/5) =
0.0892 ± 0.018 and c/5 − d/3 = −0.4306 ± 0.032 were
derived from PHOENIX stellar atmospheres, as described
in the text.
b Confidence intervals on Rp/R(cid:63) do not include the
∆Rp/R(cid:63) = 0.0006 systematic uncertainty due to stellar
variability (see text).
Table 3
Transit Parameters Inferred from Individual White Light Curves
Parameter
∆Tc (days)a
t14 (days)
R(cid:63) (R(cid:12) )
a/R(cid:63)
i (◦ )
b
Rp /R(cid:63)
Visit 3
Visit 2
Visit 1
−0.0020 0.00028 ± 0.00031 0.0002 ± 0.0004
−0.0001+0.0012
0.0357 ± 0.0007 0.0369 ± 0.0010
0.037+0.004
−0.003
0.211+0.021
0.200+0.008
0.214+0.015
−0.014
−0.006
−0.011
14.4 ± 0.9
14+1−1
15.4+0.5−0.6
89.2 ± 0.6
88.9 ± 0.7
88.5+0.8−0.7
0.21 ± 0.15
0.27 ± 0.17
0.38+0.13−0.20
0.1164+0.0009
0.1159+0.0011
0.1175+0.0011
−0.0008
−0.0009
−0.0012
predicted RMS
acheived RMS
s
337 ppm
343 ppm
1.07+0.13−0.11
337 ppm
360 ppm
1.12+0.13−0.11
337 ppm
366 ppm
1.14+0.14−0.11
a Offset between the observed mid-transit time and that cal-
culated from the linear ephemeris with P = 1.58040481 and
Tc=2454966.525123 BJDTDB .
uncertainty rescaling parameter s is slightly above but
consistent with unity, indicating the photometric scatter
is quite well explained by known sources of noise.
4.5.3. Stel lar Variability
GJ1214 is known to be variable on 50-100 day
timescales with an amplitude of 1% in the MEarth band-
pass (715-1000 nm; see Charbonneau et al. 2009; Berta
et al. 2011). To gauge the impact of stellar variability in
the wavelengths studied here, we plot in Fig. 7 the rela-
tive out-of-transit flux as measured by our WFC3 data.
For each HST visit, we have three independent measure-
ments of this quantity: the F130N narrow-band direct
image, the 0th -order spectrum, and the 1st -order spec-
trum. Consistent variability over these measurements
that sample different regions of the detector within each
visit would be difficult to reproduce by instrumental ef-
fects, such as flat-fielding errors. In Fig. 7, GJ1214 ap-
Figure 7. The relative out-of-transit (O.o.T.) flux for each HST
visit, measured independently from three different groups of im-
ages: the summed 1st -order spectrum, the 0th -order image, and
the narrow-band direct image, each normalized to its mean. Error
bars denote the standard deviation of the out-of-transit measure-
ments within each visit; they do not include the 0.5% uncertainty
in the detector flat-field. The narrow-band measurements sample
fewer photons, thus their larger uncertainties.
pears brighter in the first visit than in the last two visits,
with an overall amplitude of variation of about 1%.
This 1% variability,
if caused by unocculted spots
on the stellar surface, should lead to variations in
the inferred planet-to-star radius ratio on the order of
∆Rp/R(cid:63) = 0.0006 (Berta et al. 2011). This is larger than
the formal error on Rp/R(cid:63) from the combined white light
curve (Tab. 2), and must be considered as an important
systematic noise floor in the measurement of the abso-
lute, white-light transit depth. We do not detect this
variability in the individually measured transit depths
(Tab. 3) because it is smaller than the uncertainty on
each. Most importantly, while the spot-induced variabil-
ity influences the absolute depth at each epoch, its effect
on the relative transit depth among wavelengths will be
much smaller and not substantially bias our transmission
spectrum estimate.
4.5.4. Modeling Instrumental Systematics
Before calculating GJ1214b’s transmission spectrum,
we detour slightly to use the white light curves’s high
photometric precision to investigate the characteristics
of WFC3’s instrumental systematics. Rather than cor-
recting for the instrumental systematics with the simple
non-parametric divide-oot method, in this section we
describe them with an analytic model whose parameters
illuminate the physical processes at play. We refer to this
treatment as the model-ramp method.
In this model, we treat the systematics as consisting
of an exponential ramp, an orbit-long slope, and a visit-
(cid:16)
1 − Re−(tbat−Db )/τ (cid:17)
long slope. We relate the observed flux (Fobs ) to the
systematics-free flux (Fcor ) by
where tvis is time within a visit (= 0 at the middle of each
visit), torb is time within an orbit (= 0 at the middle of
(cid:26) D for the 1st batch of an orbit
each orbit), tbat is time within a batch (= 0 at the start
of each batch), τ is a ramp timescale, and the term
0
for the other batches
allows the exponential ramp to be delayed slightly for the
first batch of an orbit.
The exponential form arises out of the toy model pro-
posed by Agol et al. (2010), where a certain volume of
the detector pixels has the ability to temporarily trap
= (C + V tvis + B torb )
Fobs
Fcor
(5)
(6)
Db =
WFC3 Observations of GJ1214b
11
charge carriers and later release them as excess dark cur-
rent. In quick series of sufficiently strong exposures, the
population of charge traps approaches steady state, cor-
responding to the flattening of the exponential. Judging
by the appearance of the ramp in the 2nd -4th batches of
each orbit, the release timescale seems to be short enough
that the trap population completely resets to the same
baseline level after each 6 minute buffer download (dur-
ing which the detector was being continually flushed each
2.9 seconds). Compared to these batches, the 1st batch
of each orbit appears to exhibit a ramp that is either
weaker, or as we have parameterized it with the Db term,
delayed. We do not explain this, but we hypothesize that
it relates to rapid changes in the physical state of the
detector coming out of Earth occultation affecting the
pixels’ equilibrium charge trap populations.
The visit-long and orbit-long slopes are purely descrip-
tive terms (as in Brown et al. 2001; Carter et al. 2009;
Nutzman et al. 2011), but relate to physical processes in
the telescope and camera. The orbit-long slope probably
arises from the combination of pointing/focus drifts (see
Fig. 3) with our imperfect flat-fielding of the detector.
This effect of this orbital phase term could be equally
well-achieved, for instance, by including a linear func-
tion of the 0th order x and y positions (see Burke et al.
2010; Pont et al. 2007; Swain et al. 2008). The visit-long
slope is not mirrored in any of the measured geometri-
cal properties of the star on the detector, and is more
difficult to associate with a known physical cause.
In order to determine the parameters C , V , B , R, D ,
and τ , we fit Eq. 6 multiplied by a transit model to the
last three orbits of each visit’s uncorrected white light
curve. The transit parameters are allowed to vary exactly
as in §4.5.2, including the use of the informative prior on
the limb-darkening coefficients. The white light curves
with the best model-ramp fit are shown in Fig. 5, and
the properties of the residuals from this model are shown
in Fig. 6. The transit parameters from this independent
systematics correction method are consistent with those
in Tab 3. We do not quite achieve the 280 ppm predicted
scatter in the model-ramp light curves, and the residuals
show slight evidence for correlated noise (Fig. 6). More
complicated instrumental correction models could almost
certainly improve this, but we only present this simple
model for heuristic purposes. In all sections except this
one, we use the divide-oot-corrected data exclusively
for drawing scientific conclusions about GJ1214b.
Fig. 8 shows the inferred PDF’s of the instrumental
systematics parameters for all three HST visits, graphi-
cally demonstrating the striking repeatability of the sys-
tematics. As expected from the nearly identical cadence
of illumination within each of the three visits, the ramp
has the same R = 0.4% amplitude, τ = 30 second
timescale, and D = 20 second delay time across all ob-
servations. The values of τ and D are similar to the time
for a single exposure, 25 seconds (including overhead).
While the visit-long slope V is of an amplitude (fading
by 0.06% over an entire visit) that could conceivably be
consistent with stellar variability, the fact that it is iden-
tical across all three visits argues strongly in favor of
it being an instrumental systematic. B is the only pa-
rameter that shows any evidence for variability between
orbits; we would expect this to be the case if this term
arises out of flat-fielding errors, since the 1st order spec-
Figure 8. The a posteriori distribution of the instrumental sys-
tematics parameters from the analytic model, in each of the three
visits. The MCMC results for single parameters (diagonal; his-
tograms) and pairs of parameters (off-diagonal; contours encom-
passing 68% and 95% of the distribution) are shown, marginalized
over all other parameters (including c and d with priors, t14 , and
R(cid:63) ). V is measured in units of relative flux/(3 × 96 minutes), B
in relative flux/(96 minutes), R in relative flux, and both τ and
D in seconds. All visits are plotted on the same scale; for quanti-
tative comparison, the median values and 1σ uncertainties of each
parameter are quoted along the diagonal. The systematics param-
eters are remarkably repeatable from visit to visit; also, they are
largely uncorrelated with Rp /R(cid:63) (left column).
trum falls on different pixels in the three visits.
4.6. Spectroscopic Light Curve Fits
We construct multiwavelength spectroscopic light
curves by binning the extracted first order spectra into
channels that are 5 pixels (∆λ = 23 nm) wide. We esti-
mate the flux, flux uncertainty, and effective wavelength
of each bin from the inverse-variance (estimated from the
noise model) weighted average of each quantity over the
binned pixels. For each of these binned spectroscopic
light curves, we employ the divide-oot method to cor-
rect for the instrument systematics.
To measure the transmission spectrum of GJ1214b, we
fit each of these 24 spectroscopic light curves from each
of the three visits with a model in which Rp/R(cid:63) , c, d, and
s are allowed to vary. We hold the remaining parameters
fixed so that a/R(cid:63) = 14.9749 and b = 0.27729, which
are the values used by Bean et al. (2010), D´esert et al.
(2011a), and Croll et al. (2011). For limb-darkening pri-
ors, we use the same sized Gaussians on the same linear
combinations of c and d as in §4.5.1, but center them
on the PHOENIX-determined best values for each spec-
troscopic bin (see §4.2). The correlation length of all
parameters is < 10 in the MCMC chains.
For most spectroscopic bins, the inferred value of s is
within 1σ of unity, indicating that the flux residuals show
scatter commensurate with that predicted from photon
noise (1400 to 1900 ppm across wavelengths). No evi-
dence for correlated noise is seen in any of the bins, as
12
Berta et al.
judged by the same criterion as for the white light curves
(see Fig. 6).
Fig. 9 shows the transmission spectra inferred from
each of the three visits, as well as the divide-oot-
corrected, spectrophotometric light curves from which
they were derived. For the final transmission spectrum
(shown as black points in Fig. 9), we combine the three
values of Rp/R(cid:63) , and σRp /R(cid:63) in each wavelength bin by
averaging them over the visits with a weighting propor-
tional to 1/σ2
. Table 4 gives this average trans-
Rp /R(cid:63)
mission spectrum, as well as the central values of the
limb-darkening prior used in each bin. The wavelength
grids in the three visits are offset slightly (by less than a
pixel) from one another; in Table 4 we quote the average
wavelength for each bin.
In §4.5.2 we found that GJ1214’s 1% variability at
WFC3 wavelengths causes ∆D = 0.014% or ∆Rp/R(cid:63) =
0.0006 variations in the absolute transit depth. The
starspots causing this variability would have a similar
effect on measurements of the transmission spectrum,
but unless GJ1214’s starspot spectrum is maliciously be-
haved, the offsets should be broad-band and the influence
on the wavelength-to-wavelength variations within the
WFC3 transmission spectrum should be much smaller.
Each visit’s transmission spectrum is a differential mea-
surement made with respect to the integrated stellar
spectrum at each epoch; by averaging together three es-
timates to produce our final transmission spectrum, we
average over the time-variable influence of the starspots.
Importantly, if GJ1214 is host to a large population of
starspots that are symmetrically distributed around the
star and do not appear contribute to the observed flux
variability over the stellar rotation period, their effect
on the transmission spectrum will not average out (see
D´esert et al. 2011b; Carter et al. 2011; Berta et al. 2011).
If we fix the limb-darkening coefficients to the PHOENIX
values instead of using the prior, the uncertainties on the
Rp/R(cid:63) measurements decrease by 20%. If we use only
a single pair of LD coefficients (those for the white light
curve) instead of those matched to the individual wave-
length bins, the transmission spectrum changes by about
1σ on the individual bins, in the direction of showing
stronger water features and being less consistent with an
achromatic transit depth. These tests confirm that the
presence of the broad H2O feature in the stellar spectrum
(see Fig. 2) makes it especially crucial that we employ
the detailed, multiwavelength LD treatment.
As a test to probe the influence of the divide-oot sys-
tematics correction, we repeat this section’s analysis us-
ing the analytic model-ramp method to remove the in-
strumental systematics; every point in the transmission
spectrum changes by much less than 1σ . We also exper-
imented with combining the three visits’ spectroscopic
light curves and fitting for them jointly, instead of averag-
ing together the transmission spectra inferred separately
from each visit. We found the results to be practically
identical to those quoted here.
Because the transmission spectrum is conditional on
the orbital parameters we held fixed (a/R(cid:63) , b), we under-
estimate the uncertainty in the absolute values of Rp/R(cid:63) ;
the quoted σRp /R(cid:63) are intended for relative comparisons
only. Judging by the Rp/R(cid:63) uncertainty in the uncon-
strained white light curve fit (Table 2), varying a/R(cid:63)
could cause the ensemble of Rp/R(cid:63) measurements in Ta-
ble 4 to move up or down in tandem with a systematic un-
certainty that is comparable to the statistical uncertainty
on each. This is in addition to the ∆Rp/R(cid:63) = 0.0006 off-
sets expected from stellar variability (Berta et al. 2011).
4.7. Searching for Transiting Moons
Finally, we search for evidence of transiting satellite
compansions to GJ1214b in our summed WFC3 light
curves. The light curve morphology of transiting exo-
moons can be complicated, but they could generally ap-
pear in our data as shallow transit-shaped dimmings or
brightenings offset from the planet’s transit light curve
(see Kipping 2011, for a detailed discussion). While the
presence of a moon could also be detected in tempo-
ral variations of the planetary transit duration (Kipping
2009), we only poorly constrain GJ1214b’s transit dura-
tion in individual visits due to incomplete coverage.
Based on the Hill stability criterion, we would not ex-
pect moons to survive farther than 8 planetary radii away
from GJ1214b so their transits should not be offset from
GJ1214b’s by more than 25 minutes, less than the dura-
tion of an HST Earth occultation. We search only the
data in the in-transit visit, using the divide-oot method
to correct for the systematics. Owing to the long buffer
download gaps in our light curves (see §2), the most
likely indication of a transiting moon in the WFC3 light
curve would be an offset in flux from one 12-exposure
batch to another. Given the 376 ppm per-exposure scat-
ter in the divide-oot corrected light curve, we would
have expected to be able to identify transits of 0.4
R⊕ (Ganymede-sized) moons at 3σ confidence. We see
no strong evidence for such an offset. Also, we note that
starspot occultations could easily mimic the light curve
of a transiting exomoon in the time coverage we achieve
with WFC3, and such occultations are known to occur in
the GJ1214b system (see Berta et al. 2011; Carter et al.
2011; Kundurthy et al. 2011).
Due to the many possible configurations of transiting
exomoons and the large gaps in our WFC3 light curve,
our non-detection of moons does not by itself place strict
limits on the presence of exo-moons around GJ1214b.
5. DISCUSSION
The average transmission spectrum of GJ1214b from
our three HST visits is shown in Fig. 9. To the precision
afforded by the data, this transmission spectrum is flat;
a simple weighted mean of the spectrum is a good fit,
with χ2 = 20.4 for 23 degrees of freedom.
5.1. Implications for Atmospheric Compositions
We compare the WFC3 transmission spectrum to a
suite of cloud-free theoretical atmosphere models for
GJ1214b. The models were calculated in Miller-Ricci &
Fortney (2010), and we refer the reader to that paper for
their details. To compare them to our transmission spec-
trum, we bin these high-resolution (R = 1000) models to
the effective wavelengths of the 5-pixel WFC3 spectro-
scopic channels (R = 50 − 70) by integrating over each
bin. Generally, to account for the possible suppression
of transmission spectrum features caused by the overlap
of shared planetary and stellar absorption lines, this bin-
ning should be weighted by the photons detected from
WFC3 Observations of GJ1214b
13
Figure 9. Top panels: Spectroscopic transit light curves for GJ1214b, before and after the divide-oot correction, rotated and offset for
clarity. Bottom panel: The combined transmission spectrum of GJ1214b (black circles with error bars), along with the spectra measured
for each visit (colorful circles). Each light curve in the top panel is aligned to its respective wavelength bin in the panel below. Colors
denote HST visit throughout.
the system at very high resolution, but this added com-
plexity is not justified for our dataset. The normalization
of the model spectra is uncertain (i.e. the planet’s true
Rp ), so we allow a multiplicative factor in Rp/R(cid:63) to be
applied to each (giving 24-1=23 degrees of freedom for
all models). Varying the bin size between 2 and 50 pixels
wide does not significantly change any of the results we
quote in this section.
A solar composition atmosphere in thermochemical
equilibrium is a terrible fit to the WFC3 spectrum; it
has a χ2=126.2 (see Fig. 9) and is formally ruled out
at 8.2σ confidence. Likewise, the same atmosphere but
enhanced 50× in elements heavier than helium, a quali-
tative approximation to the metal enhancement in the
Solar System ice giants (enhanced 30 − 50× in C/H;
Gautier et al. 1995; Encrenaz 2005; Guillot & Gautier
2009), is ruled out at 7.5σ (χ2 = 113.2). Both models
assume equilibrium molecular abundances and the ab-
sence of high-altitude clouds; if GJ1214b has an H2 -rich
atmosphere, at least one of these assumptions would have
to be broken.
Suggesting, along these lines, that photochemistry
might deplete GJ1214b’s atmosphere of methane, D´esert
et al. (2011a), Croll et al. (2011), and Crossfield et al.
(2011) have noted their observations to be consistent
with a solar composition model in which CH4 has been
artificially removed. With the WFC3 spectrum alone,
we can rule out such an H2 -rich, CH4 -free atmosphere
at 6.1σ (Fig. 10). This is consistent with Miller-Ricci
Kempton et al. (2011)’s theoretical finding that such
thorough methane depletion cannot be achieved through
photochemical processes, even when making extreme as-
sumptions for the photoionizing UV flux from the star.
Previous spectroscopic measurements in the red opti-
cal (Bean et al. 2010, 2011) could only be reconciled with
a H2 -rich atmosphere if such an atmosphere were to host
a substantial cloud layer at an altitude above 200 mbar
(see Miller-Ricci Kempton et al. 2011). How far the flat-
tening influence of such a cloud layer would extend be-
yond 1 µm to WFC3 wavelengths would depend on both
the concentration and size distribution of the scattering
particles. As such, we explore possible cloud scenarios
consistent with the WFC3 spectrum in an ad hoc fash-
ion, using a solar composition atmosphere and arbitrarily
cutting off transmission below various pressures to emu-
late optically thick cloud decks at different altitudes in
the atmosphere. Fig. 11 summarizes the results. A cloud
deck at 100 mbar, which would be sufficient to flatten the
red optical spectrum, is ruled out at 5.7σ (χ2 = 82.8).
Due to higher opacities between 1.1 and 1.7 µm, WFC3
probes higher altitudes in the atmosphere than the red
optical, requiring clouds closer to 10 mbar to match the
data (χ2 = 23.4). Note, with the term “clouds” we refer
to all types of particles that cause broad-band extinction,
whether they scatter or absorb, and whether they were
formed through near-equilibrium condensation (such as
Earth’s water clouds) or through upper atmosphere pho-
tochemistry (such as Titan’s haze).
Fortney (2005) and Miller-Ricci Kempton et al. (2011)
identified KCl and ZnS as condensates that would be
likely to form in GJ1214b’s atmosphere, but found they
would condense deeper in the atmosphere (200-500 mbar)
than required by the WFC3 spectrum and would proba-
bly not be optically thick. While winds may be able to
14
Berta et al.
Figure 10. The WFC3 transmission spectrum of GJ1214b (black circles with error bars) compared to theoretical models (colorful lines)
with a variety of compositions. The high resolution models are shown here smoothed for clarity, but were binned over each measured
spectroscopic bin for the χ2 comparisons. The amplitude of features in the model transmission spectra increases as the mean molecular
weight decreases between a 100% water atmosphere (µ = 18) and a solar composition atmosphere (µ = 2.36).
a large mean molecular weight. We test this possibility
with H2 atmospheres that contain increasing fractions of
H2O. This is a toy model, but including molecules other
than H2 or H2O in the atmosphere would serve prin-
cipally to increase µ without substantially altering the
opacity between 1.1 and 1.7 µm, so the limits we place
on µ are robust. We find that an atmosphere with a
10% water by number (50% by mass) is disfavored by
the WFC3 spectrum at 3.1σ (χ2 = 47.8), as shown in
Fig. 9. All fractions of water above 20% (70% by mass)
are good fits to the data (χ2 < 25.5). The 10% wa-
ter atmosphere would have a minimum mean molecular
weight of µ = 3.6, which we take as a lower limit on the
atmosphere’s mean molecular weight.
For the sake of placing the WFC3 transmission spec-
trum in the context of other observations of GJ1214b, we
also display it alongside the published transmission spec-
tra from the VLT (Bean et al. 2010, 2011), CFHT (Croll
et al. 2011), Magellan (Bean et al. 2011), and Spitzer
(D´esert et al. 2011a) in Fig. 12. Stellar variability could
cause individual sets of observations to move up and
down on this plot by as much as ∆D = 0.014% for mea-
surements in the near-IR (Berta et al. 2011); we indicate
this range of potential offsets by an arrow at the right of
the plot. We display the measurements in Fig. 12 with
no relative offsets applied and note that their general
agreement is consistent with the predicted small influ-
ence of stellar variability. Depending on the temperature
contrast of the spots, however, the variability could be
larger by a factor of 2 − 3× in the optical, and we cau-
tion the reader to consider this systematic uncertainty
when comparing depths between individual studies. For
instance, the slight apparent rise in Rp/R(cid:63) toward 0.6
µm, that would potentially be consistent with Rayleigh
scattering in a low-µ atmosphere, could also be easily ex-
plained through the poorly constrained behavior of the
star in the optical. Indeed, Bean et al. (2011) found a
significant offset between datasets that overlap in wave-
length (near 0.8 µm) but were taken in different years,
suggesting variability plays a non-negligible role at these
wavelengths.
Finally, we note that any model with µ > 4, such as
one with a > 50% mass fraction of water, would be con-
sistent with the measurements from Bean et al. (2010),
D´esert et al. (2011a), Crossfield et al. (2011), Bean et al.
Figure 11. The WFC3 transmission spectrum of GJ1214b (black
circles with error bars) compared to a model solar composition
atmosphere that has thick clouds located at altitudes of 100 mbar
(pink lines) and 10 mbar red lines). We treat the hypothetical
clouds in an ad hoc fashion, simply cutting off transmission through
that atmosphere below the denoted pressures.
loft such clouds to higher altitudes, it is not clear that
the abundance of these species alone would be sufficient
to blanket the entire limb of the planet with optically
thick clouds. The condensation and complicated evo-
lution of clouds has been studied within the context of
cool stars and hot Jupiters (e.g Lodders & Fegley 2006;
Helling et al. 2008), but further study into the theoretical
landscape for equilibrium clouds on planets in GJ1214b’s
gravity and temperature regime is certainly warranted.
The scattering may also be due to a high altitude haze
formed as by-products of high-altitude photochemistry;
Miller-Ricci Kempton et al. (2011) found the conditions
on GJ1214b to allow for the formation of complex hy-
drocarbon clouds through methane photolysis.
However such clouds might form, they would either
need to be optically thick up to a well-defined altitude
or consist of a substantial distribution of particles act-
ing in the Mie regime,
i.e. with sizes approaching 1
µm. Neither the VLT spectra nor our observations give
any definitive indications of the smooth falloff in transit
depth toward longer wavelengths that would be expected
from Rayleigh scattering by molecules or small particles.
This is unlike the case of the hot Jupiter HD189733b,
where the uniform decrease in transit depth from 0.3 to
1 µm (Pont et al. 2008; Lecavelier Des Etangs et al. 2008;
Sing et al. 2011) and perhaps to as far as 3.6 µm (see Sing
et al. 2009; D´esert et al. 2009) has been convincingly at-
tributed to a small particle haze.
As an alternative,
the transmission spectrum of
GJ1214b could be flat simply because the atmosphere has
WFC3 Observations of GJ1214b
15
Table 4
GJ1214b’s Transmission Spectrum from WFC3/G141
Wavelength (µm)
1.123
1.146
1.169
1.192
1.215
1.239
1.262
1.285
1.308
1.331
1.355
1.378
1.401
1.425
1.448
1.471
1.496
1.517
1.541
1.564
1.587
1.610
1.633
1.656
Rp /R(cid:63)
0.11641 ± 0.00102
0.11707 ± 0.00099
0.11526 ± 0.00098
0.11589 ± 0.00093
0.11537 ± 0.00091
0.11574 ± 0.00090
0.11662 ± 0.00088
0.11565 ± 0.00088
0.11674 ± 0.00085
0.11595 ± 0.00087
0.11705 ± 0.00089
0.11664 ± 0.00088
0.11778 ± 0.00088
0.11693 ± 0.00091
0.11772 ± 0.00090
0.11663 ± 0.00092
0.11509 ± 0.00100
0.11635 ± 0.00104
0.11626 ± 0.00091
0.11681 ± 0.00091
0.11443 ± 0.00091
0.11631 ± 0.00091
0.11620 ± 0.00092
0.11581 ± 0.00096
c
−0.372
−0.397
−0.404
−0.410
−0.406
−0.403
−0.407
−0.403
−0.411
−0.390
−0.368
−0.371
−0.338
−0.295
−0.319
−0.322
−0.345
−0.305
−0.330
−0.351
−0.372
−0.399
−0.399
−0.415
d
1.068
1.088
1.089
1.090
1.075
1.063
1.064
1.045
1.042
1.068
1.101
1.110
1.075
1.029
1.056
1.061
1.084
1.013
1.042
1.059
1.072
1.082
1.075
1.081
(2011), and WFC3. The only observation it could not
explain would be the deep Ks -measurement from Croll
et al. (2011). Of the theoretical models we tested, we
could find none that matched all the available measure-
ments. We are uncertain of how to interpret this appar-
ent incompatibility but hopeful that future observational
and theoretical studies of the GJ1214b system may clar-
ify the issue. In the meantime, we adopt an atmosphere
with at least 50% water by mass as the most plausible
model to explain the WFC3 observations.
5.2. Implications for GJ1214b’s Internal Structure
If GJ1214b is not shrouded in achromatically optically
thick high-altitude clouds, the WFC3 transmission spec-
trum disfavors any proposed bulk composition for the
planet that relies on a substantial, unenriched, hydrogen
envelope to explain the planet’s large radius. Both the
ice-rock core with nebular H/He envelope and pure rock
core with outgassed H2 envelope scenarios explored by
Rogers & Seager (2010) would fall into this category, re-
quiring additional ingredients to match the observations.
In contrast, their model that achieves GJ1214b’s large
radius mostly from a large water-rich core, would agree
with our observations.
Perhaps most compellingly, a high µ scenario would
be consistent with composition proposed by Nettelmann
et al. (2011), who found that GJ1214b’s radius could be
explained by a bulk composition consisting of an ice-rock
core surrounded by a H/He/H2O envelope that has a wa-
ter mass fraction of 50-85%. Such a composition would
be intermediate between the H/He- and H2O-envelope
limiting cases proposed by Rogers & Seager (2010). The
H/He/H2O envelope might arise if GJ1214b had origi-
nally accreted a substantial mass of hydrogen and helium
from the primordial nebula but then was depleted of its
lightest molecules through atmospheric escape.
5.3. Prospects for GJ1214b
Future observations with the James Webb Space Tele-
scope (Deming et al. 2009; Kaltenegger & Traub 2009),
one of the immense next generation ground-based tele-
scopes (GMT, TMT, ELT; see Ehrenreich et al. 2006),
or possibly even a dedicated campaign with current fa-
cilities, could detect the 0.01% transmission spectrum
features of a 100% water atmosphere on GJ1214b, and
potentially distinguish between clear H2 -poor and cloudy
H2 -rich atmospheres. Along another front, simulations
by Menou (2011) show that observations of GJ1214b’s
thermal phase curve, such as those for HD189733b by
Knutson et al. (2007a), would probe the ratio of radiative
to advective timescales in GJ1214b’s outer envelope and
provide an independent constraint on the atmospheric
composition. Detecting the thermal emission from this
500K exoplanet is currently very difficult, and will likely
have to wait until the launch of JWST.
In the meantime, we advocate further study of the
GJ1214 system in general. Confirming and refining the
parallax for the system (van Altena et al. 1995) will im-
prove our knowledge of the stellar mass, and in turn, the
planet’s mass and radius. Likewise, further radial veloc-
ity observations will empirically constrain the hypothe-
sis by Carter et al. (2011) that a significantly non-zero
orbital eccentricity could be biasing GJ1214b’s inferred
density.
6. CONCLUSIONS
In this work, we made new measurements of the
GJ1214b’s transmission spectrum using HST/WFC3.
Reaching a precision of σRp /R(cid:63) = 0.0009 in 24 simultane-
ously measured wavelength bins, we found the transmis-
sion spectrum to be completely flat between 1.1 and 1.7
µm. We saw no evidence for the strong H2O absorption
features expected from a range of H2 -rich model atmo-
spheres.
Given the lack of a known source for clouds or hazes
that could create a truly achromatic transit depth across
all wavelengths, we interpret this flat WFC3 transmission
spectrum to be best explained by an atmosphere with a
high mean molecular weight. Based on our observations,
this atmosphere would likely consist of more than 50%
water by mass or a mean molecular weight of µ > 4. Such
an atmosphere would be consistent with observations of
GJ1214b’s transmission spectrum by Bean et al. (2010),
D´esert et al. (2011a), Crossfield et al. (2011), and Bean
et al. (2011) although it would be difficult to reconcile
with those by Croll et al. (2011).
Such a constraint on GJ1214b’s upper atmosphere
serves as a boundary condition for models of bulk com-
position and structure of the rest of the planet. It sug-
gests GJ1214b contains a substantial fraction of water
throughout the interior of the planet in order to obviate
the need for a completely H/He- or H2 -dominated enve-
lope to explain the planet’s large radius. A high bulk
volatile content would point to GJ1214b forming beyond
the snow line and migrating inward, although any such
statements about GJ1214b’s past are sub ject to large
uncertainties in the atmospheric mass loss history (see
Rogers et al. 2011).
Finally, this paper is the first published study using
WFC3 for observing a transiting exoplanet. Aside from
several instrumental systematics that were straightfor-
ward to correct and did not require a detailed instrumen-
16
Berta et al.
Figure 12. GJ1214b’s transmission spectrum from WFC3 in the context of observations from the VLT (0.6-1 µm; Bean et al. 2010, 2011),
CFHT (1.25 + 2.15 µm; Croll et al. 2011), Magellan (2.0-2.3 µm Bean et al. 2011), and Spitzer (3.6 + 4.5 µm; D´esert et al. 2011a). While
they do not measure an absolute transit depth, observations from NIRSPEC on Keck (2.1-2.3 µm Crossfield et al. 2011) disfavor models
they tested that had amplitudes larger than 0.05% in their wavelength range; we represent these constraints with the dashed rectangular
boxes. Two extremes of the models explored in this paper are shown, normalized to the MEarth-measured transit depth (see Miller-Ricci
& Fortney 2010). It is important to note that stellar variability could cause individual data sets to shift up or down on this plot as much
as ∆D = 0.014% in the near-IR or 2 − 3× more in the optical, depending on the stellar spot spectrum.
tal model, the camera delivered nearly photon-limited
performance both in individual spectrophotometric light
curves and in summed white light curves. We are con-
fident that WFC3 will serve as a valuable tool for exo-
planet atmospheric characterization in the years to come.
We thank Jacob Bean, Bryce Croll, Ian Crossfield,
Ernst de Mooij, Leslie Rogers, Drake Deming, Avi Man-
dell, Martin Kummel, Ron Gilliand, Tom Wheeler, and
the CfA Summer Statistics Club for valuable discus-
sions regarding this work. We are extremely grateful to
our Program Coordinator Patricia Royle, Contact Sci-
entist Howard Bushhouse, the WFC3 instrument team,
the entire staff at STScI and NASA, and the crew of
STS-125 for their crucial roles in enabling these obser-
vations. We thank the anonymous referee for a thor-
ough reading and thoughtful comments that improved
the manuscript. E.K. acknowledges funding from NASA
through the Sagan Fellowship Program. P.R.M. thanks
CfA director Charles Alcock for enabling a sabbatical at
SAO. We gratefully acknowledge funding from the David
and Lucile Packard Fellowship for Science and Engineer-
ing (awarded to D.C.), the National Science Founda-
tion (grant AST-0807690, awarded to D.C.), and NASA
(grant HST-GO-12251). This work is based on obser-
vations made with the NASA/ESA Hubble Space Tele-
scope, obtained at the Space Telescope Science Institute,
which is operated by the Association of Universities for
Research in Astronomy, Inc., under NASA contract NAS
5-26555. These observations are associated with program
#GO-12251. This research has made use of NASA’s As-
trophysics Data System.
Facilities: HST (WFC3)
REFERENCES
Agol, E., Cowan, N. B., Knutson, H. A., et al. 2010, ApJ, 721,
1861
Batalha, N. M., Borucki, W. J., Bryson, S. T., et al. 2011, ApJ,
729, 27
Bean, J. L., D´esert, J.-M., Kabath, P., et al. 2011, accepted to
ApJ, (arXiv:1109.0582)
Bean, J. L., Miller-Ricci Kempton, E., & Homeier, D. 2010,
Nature, 468, 669
Berta, Z. K., Charbonneau, D., Bean, J., et al. 2011, ApJ, 736, 12
Brown, T. M. 2001, ApJ, 553, 1006
Brown, T. M., Charbonneau, D., Gilliland, R. L., Noyes, R. W.,
& Burrows, A. 2001, ApJ, 552, 699
Burke, C. J., McCullough, P. R., Bergeron, L. E., et al. 2010,
ApJ, 719, 1796
Burke, C. J., McCullough, P. R., Valenti, J. A., et al. 2007, ApJ,
671, 2115
Carter, J. A., Winn, J. N., Gilliland, R., & Holman, M. J. 2009,
ApJ, 696, 241
Carter, J. A., Winn, J. N., Holman, M. J., et al. 2011, ApJ, 730,
82
Carter, J. A., Yee, J. C., Eastman, J., Gaudi, B. S., & Winn,
J. N. 2008, ApJ, 689, 499
Charbonneau, D., Berta, Z. K., Irwin, J., et al. 2009, Nature, 462,
891
Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L.
2002, ApJ, 568, 377
Charbonneau, D., Knutson, H. A., Barman, T., et al. 2008, ApJ,
686, 1341
Claret, A. 2000, A&A, 363, 1081
Claret, A., & Hauschildt, P. H. 2003, A&A, 412, 241
Croll, B., Albert, L., Jayawardhana, R., et al. 2011, ApJ, 736, 78
Crossfield, I. J. M., Barman, T., & Hansen, B. M. S. 2011, ApJ,
736, 132
Deming, D., Harrington, J., Seager, S., & Richardson, L. J. 2006,
ApJ, 644, 560
Deming, D., Seager, S., Winn, J., et al. 2009, PASP, 121, 952
D´esert, J.-M., Bean, J., Miller-Ricci Kempton, E., et al. 2011a,
ApJ, 731, L40+
D´esert, J.-M., Lecavelier des Etangs, A., H´ebrard, G., et al. 2009,
ApJ, 699, 478
D´esert, J.-M., Sing, D., Vidal-Madjar, A., et al. 2011b, A&A, 526,
A12+
Dressel, L., Wong, M., Pavlovsky, C., Long, K., et al. 2010,
”Wide Field Camera 3 Instrument Handbook”, Version 2.1,
(Baltimore: STScI)
Dunkley, J., Bucher, M., Ferreira, P. G., Moodley, K., & Skordis,
C. 2005, MNRAS, 356, 925
Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935
Ehrenreich, D., Tinetti, G., Lecavelier Des Etangs, A.,
Vidal-Madjar, A., & Selsis, F. 2006, A&A, 448, 379
Encrenaz, T. 2005, Space Sci. Rev., 116, 99
Ford, E. B. 2005, AJ, 129, 1706
Fortney, J. J. 2005, MNRAS, 364, 649
Gautier, D., Conrath, B. J., Owen, T., de Pater, I., & Atreya,
S. K. 1995, in Neptune and Triton, ed. D. P. Cruikshank,
M. S. Matthews, & A. M. Schumann, 547–611
Gibson, N. P., Pont, F., & Aigrain, S. 2011b, MNRAS, 411, 2199
Gibson, N. P., Aigrain, S., Roberts, S., et al. 2011a, accepted to
MNRAS, (arXiv:1109.3251)
WFC3 Observations of GJ1214b
17
Gregory, P. C. 2005, Bayesian Logical Data Analysis for the
Physical Sciences: A Comparative Approach with
‘Mathematica’ Support, ed. Gregory, P. C. (Cambridge
University Press)
Guillot, T., & Gautier, D. 2009, arXiv:0912.2019
Hauschildt, P. H., Allard, F., & Baron, E. 1999, ApJ, 512, 377
Helling, C., Woitke, P., & Thi, W.-F. 2008, A&A, 485, 547
Hilbert, B., & McCullough, P. 2011, WFC3 Instrument Science
Report 2011-10 (Baltimore, MD: STScI)
Hogg, D. W., Bovy, J., & Lang, D. 2010, arXiv:1008.4686
Holman, M. J., Winn, J. N., Latham, D. W., et al. 2006, ApJ,
652, 1715
Hubbard, W. B., Fortney, J. J., Lunine, J. I., et al. 2001, ApJ,
560, 413
Irwin, J. M., Quinn, S. N., Berta, Z. K., et al. 2011, accepted to
ApJ, (arXiv:1109.2055)
Kaltenegger, L., & Traub, W. A. 2009, ApJ, 698, 519
Kim Quijano, J., et al. 2009, “WFC3 Mini-Data Handbook”,
Version 1.0, (Baltimore: STScI)
Kipping, D. M. 2009, MNRAS, 392, 181
Kipping, D. M. 2011, MNRAS, 416, 689
Knutson, H. A., Charbonneau, D., Allen, L. E., et al. 2007a,
Nature, 447, 183
Knutson, H. A., Charbonneau, D., Noyes, R. W., et al. 2007b,
ApJ, 655, 564
Knutson, H. A., Madhusudhan, N., Cowan, N. B., et al. 2011,
accepted to ApJ, (arXiv:1104.2901)
Kummel, M., Kuntschner, H., Walsh, J. R., & Bushouse, H. 2011,
Instrument Science Report WFC3-2011-01 (Baltimore, MD:
STScI)
Kummel, M., Walsh, J. R., & Kuntschner, H. 2010, ”aXe User
Manual”, Version 2.1, (Baltimore: STScI)
Kummel, M., Walsh, J. R., Pirzkal, N., Kuntschner, H., &
Pasquali, A. 2009, PASP, 121, 59
Kundurthy, P., Agol, E., Becker, A. C., et al. 2011, ApJ, 731, 123
Kuntschner, H., Bushouse, H., Kummel, M., & Walsh, J. R. 2009,
ST-ECF Instrument Science Report WFC3-2009-17 (Baltimore,
MD: STScI)
Kuntschner, H., Bushouse, H., Walsh, J. R., & Kummel, M. 2008,
ST-ECF Instrument Science Report WFC3-2008-16 (Baltimore,
MD: STScI)
Kuntschner, H., Kummel, M., Walsh, J. R., & Bushouse, H. 2011,
ST-ECF Instrument Science Report WFC3-2011-05 (Baltimore,
MD: STScI)
Lecavelier Des Etangs, A., Pont, F., Vidal-Madjar, A., & Sing, D.
2008, A&A, 481, L83
L´eger, A., Rouan, D., Schneider, J., et al. 2009, A&A, 506, 287
Lissauer, J. J., Fabrycky, D. C., Ford, E. B., et al. 2011, Nature,
470, 53
Lodders, K., & Fegley, Jr., B. 2006, Chemistry of Low Mass
Substellar Ob jects, ed. Mason, J. W. (Springer Verlag), 1–+
Long, K. S., Baggett, S., Deustua, S., & Riess, A. 2010, WFC3
Instrument Science Report 2010-17 (Baltimore, MD: STScI)
Long, K. S., Wheeler, T., & Bushouse, H. 2011, WFC3
Instrument Science Report 2011-09 (Baltimore, MD: STScI)
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Markwardt, C. B. 2009, in Astronomical Society of the Pacific
Conference Series, Vol. 411, Astronomical Data Analysis
Software and Systems XVIII, ed. D. A. Bohlender, D. Durand,
& P. Dowler, 251–+
McCullough, P. 2008, Instrument Science Report WFC3 2008-26
(Baltimore, MD: STScI)
McCullough, P., & Deustua, S. 2008, Instrument Science Report
WFC3 2008-33 (Baltimore, MD: STScI)
McCullough, P. R., & MacKenty, J. W. 2011, WFC Space
Telescope Analysis Newletter 6 (Baltimore, MD: STScI)
Menou, K. 2011, submitted to ApJL, (arXiv:1109.1574)
Miller-Ricci, E., & Fortney, J. J. 2010, ApJ, 716, L74
Miller-Ricci, E., Seager, S., & Sasselov, D. 2009, ApJ, 690, 1056
Miller-Ricci Kempton, E., Zahnle, K., & Fortney, J. J. 2011,
submitted to ApJ, (arXiv:1104.5477)
Nettelmann, N., Fortney, J. J., Kramm, U., & Redmer, R. 2011,
ApJ, 733, 2
Nutzman, P., Gilliland, R. L., McCullough, P. R., et al. 2011,
ApJ, 726, 3
Pall´e, E., Zapatero Osorio, M. R., & Garc´ıa Munoz, A. 2011,
ApJ, 728, 19
Pirzkal, N., Mack, J., Dahlen, T., & Sabbi, E. 2011, Instrument
Science Report WFC3-2011-11 (Baltimore, MD: STScI)
Pont, F., Gilliland, R. L., Moutou, C., et al. 2007, A&A, 476, 1347
Pont, F., Knutson, H., Gilliland, R. L., Moutou, C., &
Charbonneau, D. 2008, MNRAS, 385, 109
Rogers, L. A., Bodenheimer, P., Lissauer, J. J., & Seager, S. 2011,
ApJ, 738, 59
Rogers, L. A., & Seager, S. 2010, ApJ, 716, 1208
Sada, P. V., Deming, D., Jackson, B., et al. 2010, ApJ, 720, L215
Seager, S., & Sasselov, D. D. 2000, ApJ, 537, 916
Seager, S., Kuchner, M., Hier-Ma jumder, C. A., & Militzer, B.
2007, ApJ, 669, 1279
Sing, D. K., D´esert, J.-M., Lecavelier Des Etangs, A., et al. 2009,
A&A, 505, 891
Sing, D. K., Pont, F., Aigrain, S., et al. 2011, MNRAS, 1159
Sivia, D. S. 1996, Data Analysis: A Bayesian Tutorial (Oxford
University Press), xi, 189 p.
Smith, R. M., Zavodny, M., Rahmer, G., & Bonati, M. 2008a, in
Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference Series, Vol. 7021, Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series
Smith, R. M., Zavodny, M., Rahmer, G., & Bonati, M. 2008b, in
Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference Series, Vol. 7021, Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series
Southworth, J. 2008, MNRAS, 386, 1644
Swain, M. R., Vasisht, G., & Tinetti, G. 2008, Nature, 452, 329
van Altena, W. F., Lee, J. T., & Hoffleit, E. D. 1995, The general
catalogue of trigonometric [stellar] parallaxes, ed. van Altena,
W. F., Lee, J. T., & Hoffleit, E. D.
van Hamme, W. 1993, AJ, 106, 2096
Viana, A. C., & Baggett, S. 2010, WFC3 TV3 Testing: IR
Crosstalk, Tech. rep.
Winn, J. N., Matthews, J. M., Dawson, R. I., et al. 2011, ApJ,
737, L18+
|
1801.07707 | 4 | 1801 | 2018-07-30T13:33:52 | Pebble dynamics and accretion onto rocky planets. I. Adiabatic and convective models | [
"astro-ph.EP"
] | We present nested-grid, high-resolution hydrodynamic simulations of gas and particle dynamics in the vicinity of Mars- to Earth-mass planetary embryos. The simulations extend from the surface of the embryos to a few vertical disk scale heights, with \rev{a spatial} dynamic range \rev{of} $\sim\! 1.4\times 10^5$. Our results confirm that "pebble"-sized particles are readily accreted, with accretion rates continuing to increase up to metre-size "boulders" for a 10\% MMSN surface density model. The gas mass flux in and out of the Hill sphere is consistent with the Hill rate, $\Sigma\Omega R_\mathrm{H}^2 = 4\, 10^{-3}$ M$_\oplus$ yr$^{-1}$. While smaller size particles mainly track the gas, a net accretion rate of $\approx 2\,10^{-5}$ M$_\oplus$ yr$^{-1}$ is reached for 0.3--1 cm particles, even though a significant fraction leaves the Hill sphere again. Effectively all pebble-sized particles that cross the Bondi sphere are accreted. The resolution of these simulations is sufficient to resolve accretion-driven convection. Convection driven by a nominal accretion rate of $10^{-6}$ M$_\oplus$ yr$^{-1}$ does not significantly alter the pebble accretion rate. We find that, due to cancellation effects, accretion rates of pebble-sized particles are nearly independent of disk surface density. As a result, we can estimate accurate growth times for specified particle sizes. For 0.3--1 cm size particles, the growth time from a small seed is $\sim$0.15 million years for an Earth mass planet at 1 AU and $\sim$0.1 million years for a Mars mass planet at 1.5 AU. | astro-ph.EP | astro-ph | MNRAS 000, 1 -- 21 (2018)
Preprint 31 July 2018
Compiled using MNRAS LATEX style file v3.0
Pebble dynamics and accretion onto rocky planets.
I. Adiabatic and convective models
Andrius Popovas,1,3(cid:63) Ake Nordlund,1† Jon P. Ramsey,1‡ and Chris W. Ormel,2
1Center for Star and Planet Formation, the Niels Bohr Institute and the Natural History Museum of Denmark,
University of Copenhagen, Øster Voldgade 5-7, DK-1350 Copenhagen, Denmark
2Anton Pannekoek Institute, University of Amsterdam, Science Park 904, C4.149, 1098 XH Amsterdam, The Netherlands,
3Rosseland Centre for Solar Physics, Institute of Theoretical Astrophysics, University of Oslo, P.O. Box 1029 Blindern, N-0315 Oslo, Norway
Accepted 2018 June 27. Received 2018 June 14; in original form 2018 January 26
ABSTRACT
We present nested-grid, high-resolution hydrodynamic simulations of gas and particle
dynamics in the vicinity of Mars- to Earth-mass planetary embryos. The simulations
extend from the surface of the embryos to a few vertical disk scale heights, with a
spatial dynamic range of ∼ 1.4 × 105. Our results confirm that "pebble"-sized parti-
cles are readily accreted, with accretion rates continuing to increase up to metre-size
"boulders" for a 10% MMSN surface density model. The gas mass flux in and out
H = 4 10−3 M⊕ yr−1. While
of the Hill sphere is consistent with the Hill rate, ΣΩR2
smaller size particles mainly track the gas, a net accretion rate of ≈ 2 10−5 M⊕ yr−1 is
reached for 0.3 -- 1 cm particles, even though a significant fraction leaves the Hill sphere
again. Effectively all pebble-sized particles that cross the Bondi sphere are accreted.
The resolution of these simulations is sufficient to resolve accretion-driven convection.
Convection driven by a nominal accretion rate of 10−6 M⊕ yr−1 does not significantly
alter the pebble accretion rate. We find that, due to cancellation effects, accretion rates
of pebble-sized particles are nearly independent of disk surface density. As a result,
we can estimate accurate growth times for specified particle sizes. For 0.3 -- 1 cm size
particles, the growth time from a small seed is ∼0.15 million years for an Earth mass
planet at 1 AU and ∼0.1 million years for a Mars mass planet at 1.5 AU.
Key words: planets and satellites: formation, terrestrial planets -- protoplanetary
discs -- hydrodynamics (HD) -- methods: numerical
1 INTRODUCTION
We now know that the majority of stars in our galaxy host
at least one planet (Cassan et al. 2012). We also know that
planets form inside the gaseous and dusty protoplanetary
disks (PPDs) that orbit stars during their youth. As it is now
possible to spatially resolve the structure of young PPDs
(e.g., with the Atacama Large Millimetre/sub-millimetre
Array; ALMA), it is becoming apparent that the major-
ity of PPDs have a rich sub-structure of rings, gaps, vor-
tices and spirals (e.g. van der Marel et al. 2013; Isella et al.
2013; Casassus et al. 2013; ALMA Partnership et al. 2015;
Andrews et al. 2016; Meru et al. 2017). Recent, spatially
resolved observations of young disks with ALMA (ALMA
Partnership et al. 2015; Andrews et al. 2016), as well as
meteoritic evidence (Bizzarro et al. 2017), furthermore in-
(cid:63) E-mail: [email protected] (AP)
† E-mail: [email protected] (AN)
‡ E-mail: [email protected] (JPR)
© 2018 The Authors
dicate that planet formation starts early. What remains to
be answered, however, is how planets grow efficiently enough
under realistic conditions to agree with the astronomical ob-
servations and meteoritic evidence.
There are currently two scenarios for planet growth
which are popular in the literature: First, the planetesimal
accretion scenario, in which a planetary embryo is impacted
by kilometre-size bodies and their mass is added to the em-
bryo (e.g. Pollack et al. 1996; Hubickyj et al. 2005). Even-
tually, the embryo becomes massive enough that gravita-
tional focusing (Greenberg et al. 1978) becomes important
and the embryo enters what is commonly called the 'runaway
growth' phase. Planetesimal accretion ends when there is no
remaining solid material in an embryo's 'feeding zone', e.g.,
when the embryo grows massive enough to gravitationally
stir up nearby planetesimals, enhancing their eccentricities
and effectively kicking them out of the feeding zone (Ida
& Makino 1993). The growth of the, now somewhat iso-
lated, protoplanet transitions to the much slower 'oligarchic
growth' phase (Kokubo & Ida 1998). In this phase, the
8
1
0
2
l
u
J
0
3
.
]
P
E
h
p
-
o
r
t
s
a
[
4
v
7
0
7
7
0
.
1
0
8
1
:
v
i
X
r
a
2
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
few, largest mass protoplanets grow oligarchically, while the
remaining planetesimals mostly remain small. The critical
time scale in this context is the lifetime of the PPD, which
is of the order of several million years (e.g. Bell et al. 2013).
However, numerical simulations of the planetesimal accre-
tion hypothesis have the problem that they predict that it
takes much longer than a PPD lifetime for cores to grow to
observed planetary masses (e.g. Levison et al. 2010; Bitsch
et al. 2015b).
K , where ΩK =(cid:112)GM∗/r3 is the Keplerian frequency,
The "pebble accretion" scenario has, meanwhile, demon-
strated promise in being able to accelerate the protoplanet
growth process significantly (Ormel & Klahr 2010; Johansen
& Lacerda 2010; Nordlund 2011; Lambrechts & Johansen
2012; Morbidelli & Nesvorny 2012; Lambrechts & Johansen
2014; Bitsch et al. 2015b; Chatterjee & Tan 2014; Visser &
Ormel 2016; Ormel et al. 2017, etc.). "Pebbles", in the astro-
physical context, are millimetre to centimetre-sized particles
with stopping times, ts, comparable to their orbital period,
ts ∼ Ω−1
G is the gravitational constant, M∗ and r are the mass of
and the distance to the central star. Pebbles are likely to
form a significant part of the solid mass budget in PPDs, as
indicated by both dust continuum observations (e.g. Testi
et al. 2003; Lommen et al. 2009) and by the mass frac-
tion of chondrules in chondritic meteorites (Johansen et al.
2015; Bollard et al. 2017). In this scenario, the combined
effects of gas drag due to the difference in speed between
the slightly sub-Keplerian dust (and gas) and the Keplerian
embryo, and gravity ensure that pebbles are captured and
settle to the planet. As the embryo grows, the efficiency of
gravitational focusing of pebbles increases, and the effective
accretion cross section becomes larger than the embryo it-
self.
The radius of dominance of the gravitational force of a
planet relative to the central star is approximately given by
the Hill radius:
(cid:115)
RH = a 3
Mp
3M∗ ,
(1)
where a is the semi-major axis of the embryo's orbit, Mp
and M∗ are the masses of the embryo and the central star,
respectively. Particles of suitable size, passing the embryo
even as far away as the Hill radius may be accreted, as has
been shown analytically by Ormel & Klahr (2010), using test
particle integrations on top of hydrodynamical simulations
by Morbidelli & Nesvorny (2012) and using numerical sim-
ulations with particles by Lambrechts & Johansen (2012).
According to these works, pebble accretion is so efficient
that it can reduce planet growth time scales to well within
the lifetime of a PPD, even at large orbital distances from
the host star.
Herein, we present the first-ever high resolution simula-
tions of gas and pebble dynamics in the vicinity of low-mass,
"rocky" planetary embryos (for simplicity, often referred to
only as 'embryos' in what follows) and report on the mea-
sured accretion efficiency of pebbles. We consider three em-
bryo masses: 0.95 M⊕, 0.5 M⊕, and 0.096 M⊕, and we specif-
ically choose to study conditions expected for pressure traps
associated with inside-out-scenarios of planet formation (e.g.
Tan et al. 2016). The embryos are thus assumed to be em-
bedded in disks that have a local pressure maximum at the
orbital radius of the embryo.
The presence of a pressure bump instead of a flat density
distribution provides three advantages:
(i) It eliminates the head wind otherwise associated with
the slightly sub-Keplerian motion of the gas;
(ii) It helps to trap particles drifting inwards from larger
orbital radii (Whipple 1972; Paardekooper & Mellema 2006);
(iii) High resolution observations (e.g. ALMA Partner-
ship et al. 2015; Andrews et al. 2016) show disks with rings
and gaps, which affect the distribution of solids in the disk.
Having the pressure bump allows particles to drift into the
pressure maximum, thus mimicking the observed behaviour.
The structure of the paper is as follows. In Section 2 we
describe the simulation set up, including initial conditions
and their dynamic relaxation to a quasi-stationary state.
We then describe, in Section 3, our treatment of particles,
their equations of motion and how they are injected into
the simulation domain. In Section 4 we describe the gas dy-
namics that develops in our simulations and compare our
results to other studies. In Section 5 we present our main
results regarding particle dynamics and accretion efficiency
onto planetary embryos. In Section 6 we present the first
results of simulations of accretion-driven convection in the
primordial atmospheres of rocky planets. Finally, in Section
7, we summarize our main results, discuss their implications
and indicate future work.
2 SIMULATION SET UP
This study is carried out using the new DISPATCH frame-
work (Nordlund et al. 2018) in a three-dimensional, Carte-
sian (shearing box) domain, with a set of static, nested
patches. We employ a variation of the finite-difference
STAGGER solver (Nordlund et al. 1994; Kritsuk et al. 2011)
that, apart from the effects of transformation to a rotating
coordinate system detailed below, solves the following set of
partial differential equations:
= −∇ · (ρu);
= −∇ · (ρuu + τ
= −∇ · (su),
=
∂ ρ
∂t
∂ ρu
∂t
∂s
∂t
) − ∇(p + pa) − ρ∇Φ;
(2)
(3)
(4)
on a staggered mesh, where ρ is the gas density, u is
the gas velocity, p is the gas pressure, pa is a stabilis-
ing artificial pressure term, τ
is a viscous stress tensor,
=
s = ρ log(p/ργ)/(γ − 1) is the entropy density per unit vol-
ume, γ is the ratio of specific heats and Φ is the gravitational
potential. The viscous stress tensor is taken to be
= τi, j = ν1∆x(cs + u)ρ
(cid:18) ∂ui
(cid:19)
(5)
+
∂uj
∂xi
,
∂xj
τ
=
where ∆x is the grid resolution, cs is the sound speed, and
the coefficient ν1 is a small fraction of unity. The artificial
pressure is computed as:
pa = ν2 ρ min(0, ∆x(∇ · u))2,
where ν2 is a coefficient ∼ 1 (Kritsuk et al. 2011).
(6)
This allows a discretisation that explicitly conserves
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
3
mass, momentum and entropy. The viscous stress tensor sta-
bilizes the code in general, while the artificial pressure term
ensures that shocks are marginally resolved.
The equation of state (EOS) considered herein is that
of an ideal gas with adiabatic index γ = 1.4 and molecular
weight µ = 2. Although not considered here, in a forthcoming
paper we include radiative energy transfer and adopt a more
realistic, table-based EOS (Popovas et al., in prep.).
We use code units where length is measured in Earth
radii, mass is measured in Earth masses, and velocity is mea-
sured in units of 1 km s−1. Hence our time unit is R⊕/1km s−1
= 6371 seconds, corresponding to roughly 5,000 time units
per year.
Partly because the time step is determined locally in
each patch, and partly because of the small number of cells
per patch (333 in this case), the hydro solver operates near
its optimal speed -- updating on the order of 1.4 million cells
per core per second -- we performed these simulations using
only a single, 20-core Intel Ivy Bridge node for each experi-
ment.
2.1 Grid set up
We use nested sets of Cartesian patches centred on the em-
bryo. With the embryo as the origin of our coordinate sys-
tem, the x-axis is parallel to the cylindrical radial direction
and points away from the central star, while the y- and z-
axes are oriented along the azimuthal and vertical directions,
respectively. The +y-axis lies along the direction of orbital
rotation. The embryo is assumed to move in a circular or-
bit. Patches are arranged in a nested, hierarchical 'Rubik's
Cube' arrangement,with 216 patches arranged into a 6×6×6
cube. The central 8 patches are then split recursively into
6 × 6 × 6 child patches, which have a three times finer reso-
lution than the previous level. This is repeated for up to 7
levels of refinement (8 levels for the smallest embryo mass),
while the coarsest set is duplicated 5 times in the y-direction.
The size of the innermost, central, patch, exceeds the em-
bryo radius RP by 10-30%. With 7 levels in the hierarchy,
the outer scale in the vertical and radial directions is then
∼ 5000RP, which is about 4 times a typical disk scale height
for an Earth-mass planet, and ∼20x larger than the Hill ra-
dius. In the y-direction, the model extends to ∼ 25000RP.
Each patch contains 33×33×33 cells, leading to a maximum
resolution in our simulations, near the embryo's surface, of
≈ 3%RP. This resolution is higher than has been used in pre-
vious works (e.g. D'Angelo & Bodenheimer 2013; Cimerman
et al. 2017; Lambrechts & Lega 2017). Since the numerical
viscosity is proportional to the cell size (cf. Eq. 5), the nu-
merical viscosity in our simulations is also correspondingly
smaller than in these previous studies1.
2.2 Initial conditions
We consider a local shearing flow in the reference frame that
co-moves with the embryo at the local Keplerian frequency,
ΩK. We ignore the slight curvature of the partial orbit but
include the Coriolis force resulting from the orbital motion.
The origin of the co-moving frame is, depending on the simu-
lation run, placed at either 1 or 1.5 astronomical units (AU).
The initial gas velocity is then given by:
u(t = 0) = −
(cid:18) 3
(cid:19)
(7)
+ ζp
ΩKxy,
2
where ζp is a pressure trap parameter, defined below. In
this study, we do not consider a systematic accretion flow
towards the central star, nor the headwind that an embryo
experiences when the gas is slightly sub-Keplerian. In ana-
lytically constructed accretion disks, with pressure decreas-
ing systematically with radius, the headwind is typically on
the order of one percent of the Keplerian velocity, but when
an embryo is embedded in a local pressure bump there is
no headwind. A systematic accretion flow would allow the
smallest size particles to escape the pressure trap, and would
thus tend to increase the fraction of large size particles, but
since we are mainly concerned with measuring the accretion
efficiency of different size particles, the possible evolution of
the grain size distribution is not of direct importance for our
study.
The background vertical density stratification is set by
the balance between the stellar gravity and the gradient of
the gas pressure:
(cid:32)
(cid:33)
ρ(z) = ρ0 exp
− Ω2
Kz2 + 2ζp x2
2p0/ρ0
,
(8)
where ρ0 and p0 are the disk midplane density and pres-
sure, respectively, and are estimated from the 2D radiative
disk models of Bitsch et al. (2015a). The ζp term gives rise
to a pressure gradient that balances the extra shear intro-
duced in Eq. 7 above. A pressure bump produces a more
self-contained region in the vicinity of the embryo's orbit,
which retains the density wake produced by the embryo,
preventing it from causing a forceful numerical bounce at
the boundaries. This is also mitigated by the wave-killing
boundaries that we use (see below), but a less forceful wake
means a smaller region at the boundaries needs to be artifi-
cially damped.
With typical dust opacities2 on the order of 0.1 - 1 cm2
g−1, disks with surface densities similar to the Minimum
Mass Solar Nebula (MMSN; Hayashi 1981) are optically
thick in the vertical direction, since the MMSN nominally
has 1700 g/cm2 at 1 AU. If early hydrostatic atmospheres
that surround planetary embryos are assumed to be nearly
adiabatic one can predict their approximate vertical struc-
ture from knowledge of only the density, ρ0, and tempera-
ture, T0, at the outer boundary. The pressure at the outer
boundary of the atmosphere is assumed to match the local
pressure of the surrounding disk. Therefore, for reasonable
initial conditions, we approximate an adiabatic atmosphere
profile outside-in to the embryo. From the EOS we have the
1 As demonstrated in Kritsuk et al. (2011), the effects of numer-
ical viscosity are generally not larger in codes that use explicit
numerical viscosity than in codes that rely on viscous effects im-
plicit to the methods.
2 Chondritic meteorites typically have a fair fraction of "matrix",
which corresponds to essentially unprocessed dust. This demon-
strates that, at least during the epoch when chondritic parent
bodies formed, dust provided substantial opacity per unit mass
in the proto-Solar disk.
MNRAS 000, 1 -- 21 (2018)
4
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
and the Bondi radius,
,
RB =
GMp
c2
iso
where c2
iso is the isothermal sound speed. In fluid dynamics,
the Bondi radius has no particular dynamical significance,
other than being the radius where the pressure scale height
of the atmosphere becomes comparable to the distance from
the embryo, and the radius where trans-sonic bulk gas speeds
would become unbound.
(13)
Note that the mass in such an adiabatic atmosphere
is completely determined (apart from the equation of state)
by the pressure and temperature at the outer boundary, and
that if the external pressure is reduced the resulting mass is
smaller. This implies that mass must be lost from such atmo-
spheres when the disk evolves to lower densities (Nordlund
2011; Schlichting et al. 2015; Ginzburg et al. 2016; Ruba-
nenko et al. 2017).
2.2.1 Parameter space
In this study, we consider 8 cases: 3 different embryo masses
(Mp = {0.95, 0.5, 0.096} M⊕) at 2 distances from the central
star (a = {1.0, 1.5} AU), with different disk midplane densi-
ties (ρ0 = {1 10−10, 4 10−11} g cm−3) and temperatures (T0
= {230, 300} K), as well as with/without a pressure bump
(ζp = {0.0, 0.05, 0.1}). The basic parameters for all cases
are summarised in Table 1.
The radii of the embryos are estimated using the mass-
radius formula from Zeng et al. (2016):
= (1.07 − 0.21 CMF)
M⊕
R⊕
where CMF stands for core mass fraction. We take CMF =
0.325, although the exact value is not important; taking
CMF = 0.0 only changes the radius of the embryo on the
order of 5%.
(14)
(cid:18) M
(cid:19)1/3.7
,
(cid:18) RP
(cid:19)
Figure 1. Radial temperature profile for a planetary embryo at-
mosphere using an ideal gas EOS (γ = 1.4, blue curve) and a
realistic, tabular EOS (red curve).
pressure at the outer boundary:
P0 =
ρ0kBT0
µmu
,
(9)
where kB is the Boltzmann constant and mu is the atomic
mass unit. For each layer of the atmosphere, ri, we iterate
over the set of equations:
,
∆ ln Ti = ∆ ln Pi
(12)
where ∆ ln ri = 2(ri+1 − ri)/(ri+1 + ri). The temperature and
pressure for each layer are then simply Ti = Ti+1e∆ ln Ti and
Pi = Pi+1e∆ ln Pi , respectively. We use 500 logarithmically-
spaced layers between half the embryo radius and the Hill
radius. We then do a polynomial interpolation between these
layers to map the density, pressure and temperature to each
cell in the simulation box in order to construct the initial
conditions in the vicinity of the embryo.
ρi =
Pi µmu
kBTi
∆ ln Pi = ∆ ln ri
(cid:18)
;
GMp
2
γ − 1
γ
(cid:19)
;
(10)
(11)
ρi+1
Pi+1ri+1
+
ρi
Piri
For the purposes of this study, an ideal gas EOS (γ =
1.4) is sufficient. As demonstrated by Fig. 1, the thermal
profiles of the atmosphere surrounding an M = 0.95M⊕ em-
bryo given by an ideal gas EOS and a realistic, tabular EOS
(Tomida et al. 2013, updated by Tomida & Hori 2016) are
only significantly different close to the surface of the embryo.
Gas flows in the vicinity of the Hill sphere, as well as embryo
particle accretion rates, are insensitive to these differences.
With constant opacities and an internal heating source
(due to the accretion of solids) such an atmosphere is convec-
tive, and hence has a nearly adiabatic structure (Stevenson
1982). The atmosphere might have a radiative outer shell --
cf. the works by Piso et al. (2015) on giant planets and by In-
aba & Ikoma (2003) for a terrestrial mass planet -- but this
shell would not substantially change the physical structure
of the atmosphere, and ignoring it is a reasonable approx-
imation when constructing the initial conditions. In forth-
coming work, however, we include radiative energy transfer
in our simulations, and there the initial conditions take this
radiative shell into account.
The effective, dynamical boundary between the disk and
the embryo atmosphere occurs at distances from the plane-
tary embryo somewhere in the range between the Hill radius
2.3 External forces (accelerations)
r3
+ 3Ω2
= − GMp(xx + yy + zz)
External forces, arising from the transformation to a rotat-
ing coordinate system, and from linearising the force of grav-
ity due to the central star, are added to the hydrodynamics
as source terms:
∂u
Kzz − 2ΩKz× u, (15)
∂t
where r is the distance from the embryo. The first term in
Eq. 15 is the two-body force from the embryo. The second
term is the tidal force and has two contributions: the cen-
trifugal force and the differential change of the stellar grav-
ity with radius. Vertically, there is a contribution from the
projection of the force of gravity onto the z-axis (the third
term). The last term is the classical Coriolis force.
Kxx − Ω2
To avoid the singularity of the two-body force, rather
than using a 'softening' of the potential (e.g. Ormel et al.
2015a; Lambrechts & Lega 2017; Xu et al. 2017), we truncate
it via r = max(r, 0.1RP). Truncation is better than softening,
since the gas outside the radius of the embryo feels the cor-
rect force, and the spherical boundary conditions (described
below) negates the effect of truncation for the innermost
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
5
Table 1. Basic parameters for the different simulation runs. See the text for more details.
Run
m095t00
m095t05
m095t10
m05t00
m05t05
m05t10
m01t00
m01t10
m095t10-conv
Mp [M⊕]
Rp [R⊕]
a [AU]
T0 [K]
0.95
0.95
0.95
0.5
0.5
0.5
0.096
0.096
0.95
0.988
0.988
0.988
0.83
0.83
0.83
0.532
0.532
0.988
1.0
1.0
1.0
1.0
1.0
1.0
1.5
1.5
1.0
300
300
300
300
300
300
230
230
300
ρ0 [g cm−3]
1 10−10
1 10−10
1 10−10
1 10−10
1 10−10
1 10−10
4 10−11
4 10−11
1 10−10
ΩK
2π [yr−1]
ζp
xdamp
∆xmin [R⊕]
box [R⊕]
1.0
1.0
1.0
1.0
1.0
1.0
0.5433
0.5433
1.0
0.0
0.05
0.1
0.0
0.05
0.1
0.0
0.1
0.1
700
700
700
500
500
500
400
400
700
0.037
0.037
0.037
0.031
0.031
0.031
0.023
0.023
0.037
5186 x 5186 x 25930
5186 x 5186 x 25930
5186 x 5186 x 25930
4360 x 4360 x 21801
4360 x 4360 x 21801
4360 x 4360 x 21801
9773 x 9773 x 48866
9773 x 9773 x 48866
5186 x 5186 x 25930
cells, well within the radius of the spherical boundary con-
dition/embryo.
2.4 Boundary conditions
Using test simulations, we have ensured that our boundary
conditions remain stable over long periods of integration
time and under rather extreme conditions (e.g. highly su-
personic flows, low densities). The "azimuthal" y-boundaries
are set to be periodic, whereas the vertical z- and "radial"
x-boundaries are set as follows:
• Zero-gradient conditions for the entropy per unit mass.
• The x- and z-velocity components are set symmetric
with respect to the boundaries, while the y-component is
extrapolated in the x-direction and z directions.
• The density boundary conditions in the x-direction is
constructed from assuming hydrostatic balance:
ρx = ρxi exp
− Ω2
Kζp
T0
(x2 − x2
i )
,
(16)
(cid:32)
(cid:32)
(cid:33)
(cid:33)
where ρx is the density in the boundary at location x and
ρxi is the density at location xi, which corresponds to the
last cell inside a patch's active domain.
• If a pressure trap is not present we assume a vanishing
• In the vertical direction, we apply a similar condition:
derivative of the density in the x-direction.
ρz = ρz0 exp
− Ω2
K
2T0
(z2 − z2
0)
,
(17)
where ρz is the density in the boundary at location z and
ρz0 is the density at location z0, which again corresponds to
the last cell inside the active domain.
We additionally apply the "wave-killing" prescription of
de Val-Borro et al. (2006) to reduce reflections at the x-
boundaries. It is only applied over a specific range, given by
xdamp (column 9 in Table 1).
2.4.1 Spherical boundaries
In these simulations we use "jagged", no-slip boundaries to
approximate the embryo surface. Fig. 2 shows a 2D repre-
sentation of these boundary conditions on a staggered mesh.
Density and entropy are allowed to evolve freely everywhere;
only the mass flux is constrained. We deem a set of cell
MNRAS 000, 1 -- 21 (2018)
Figure 2. Treatment of spher-
ical boundary conditions on
a staggered mesh. Blue and
red squares represent horizon-
tal and vertical mass fluxes
outside the boundary, which
evolve freely, whereas turquoise
and magenta squares represent
mass fluxes inside the bound-
ary, which are set
to zero.
Green squares represent den-
sity points that are subject to
change, while yellow squares
show density points that re-
main constant throughout the
simulation.
faces that approximate the spherical surface to belong to the
boundary. If the cell face is inside the spherical boundary,
the mass flux is set to zero, otherwise it is allowed to evolve
freely. Hence, density cannot change inside the boundary
and cells that have faces belonging to the boundary react
exactly as they should: inflows towards the boundary pro-
duce a positive ∂ ρ/∂t at the boundary. The corresponding
increase in pressure causes an increase in the outward di-
rected pressure gradient at the last 'live' mass flux point, as
expected at a solid boundary. This permits the system to
find an equilibrium and react to perturbations from it, and
since there are never any inflows and outflows through the
boundary, mass is conserved.
2.5 Relaxation of initial conditions
We allow the system to relax for 30 000 time units (which
corresponds to about 6 years), in order to ensure that dy-
namically consistent conditions are attained. Indeed, as dis-
cussed below, tests demonstrate that it takes on the order
of only one orbit for the initial conditions to relax.
The density stratification well inside the Hill sphere is
almost hydrostatic, but with deviations and fluctuations on
short time scales, and it is thus not necessary to include
the detailed dynamics of this region while larger structures
are relaxing. Therefore, we begin our experiments by fixing
the hydrostatic structure there, and set the spherical bound-
ary to a radius of 27 embryo radii. We then 'release' layers
(add an extra level of refinement, reduce the radius of the
6
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
spherical boundary, turn on the dynamics outside of the new
spherical boundary) at successively smaller radii (9, 3 and
finally, 1 embryo radii). The 'releases' are typically done at
times = 10 000, 20 000 and 26 000 in code units. Successive
releases are increasingly more expensive due to the smaller
cell size and larger sound speeds, but need less time in code
units to adjust to the surrounding environment (for essen-
tially the same reasons, i.e., smaller scales and larger sound
speeds). This procedure saves significant amounts of com-
puting time during the initial relaxation phase.
After the last release, the simulations are allowed to
relax for an additional 4000 time units before they are
deemed ready for the next phase. Figures 3 and 4 show large-
scale and zoomed-in density slices at the midplane for run
m095t00 at t = 0 and after a relaxed state is reached at t =
30 000. Panels on the right show the relaxed state with well
established horseshoe orbits, density wakes, and slight den-
sity perturbations ahead/behind of the embryo at the same
orbital distance (discussed in more detailed below).
Fig. 4 is a zoom-in of the same run with the Bondi ra-
dius more clearly visible (largest red circle); the Rubik's cube
nesting of grids is also visible. The dashed lines indicate the
successive releases of the radius of the embryo (green - 27,
yellow - 9, light red - 3 embryo radii). The tiny black circle
corresponds to one embryo radii. The right panel shows the
simulation when fully relaxed and, via streamlines, demon-
strates that the disk gas penetrates deeply into RB, down
to ≈ 20RP. At even smaller radii the atmosphere is nearly
static, with motions that have a weak rotational tendency
(see below).
Fig. 5 shows the physical structure of the primordial
atmosphere at both time = 0 and 30 000. The cuts are done
along the x-axis in the midplane and along the z-axis at
x = y = 0. As can be seen, the initial state is very close to
the final state of the atmosphere, and there are only negli-
gible variations for different cuts. Although the density and
temperature profiles have changed slightly, the changes are
minimal and the use of these initial conditions thus saves a
lot of relaxation time (compared to, e.g., a sudden insertion
of the planet's gravitational potential, which would require
waiting tens of orbits with the maximum level of refinement
for the dynamics to relax). Although the indnermost parts
of the atmosphere still retain some weak random motions,
by time = 30 000, the run has nearly reached a steady-
state, when injection of particles can commence. Note that,
as has been pointed out by Ormel et al. (2015b), a com-
pletely steady, static state will likely never be reached in the
vicinity of the embryo.
3 PARTICLES
After relaxation of the fluid motions and density distribution
for 30,000 time units (about 6 years) we add particles for the
purpose of tracing the motions of selected sub-samples and
reveal, for example, the typical paths of different physical
size particles and to be able to measure the rate of accre-
tion onto the planetary embryos. In these simulations, we
consider macro-particles (sometimes called super-particles;
Rein et al. 2010; Ebisuzaki & Imaeda 2017), each of which
represents a swarm of real particles, with a given size and
mass. We follow the particle motions, taking into account ex-
ternal forces and gas drag, until they are accreted onto the
planet embryo, or exit the domain through the x-boundaries.
We consider particles as 'accreted' when their radial distance
to the centre of the planet is smaller than the radius of the
embryo.
In order to have statistical coverage everywhere, the
initial spatial distribution of macro-particles is chosen to be
proportional to the local gas density. The actual dust-to-gas
ratio is likely characterized by a strong settling towards the
midplane, while the dust-to-gas surface density ratio should
be essentially constant in the upstream flow entering the Hill
sphere. However, rather than having to make assumptions
about the settling, we instead analyse sub-populations of our
initial distribution, tagging and following for example only
the particles that initially reside within a given distance from
the midplane.
We use macro-particles with corresponding real particle
sizes normally ranging from 10 µm to 1 cm, with a constant
number of macro-particles per logarithmic size bin. However,
note that we never integrate over size. The constant number
per bin was chosen to provide good statistical coverage for all
sizes when investigating the transport of particles in narrow
size bins.
We include 1 cm particles for completeness, even though
such large particles are typically found only in CB chondrites
(Friedrich et al. 2015), partly because there is no guarantee
that the size distribution of chondrules in chondrites spans
the entire size range that was available at the time of pebble
accretion onto planet embryos. We also made complemen-
tary experiments with even larger size particles (up to 1 m),
to explore saturation of accretion rates with size.
The shape of the relative size distribution will be mod-
ified near the embryo, because large particles accrete more
rapidly onto the embryo, while being replenished at the same
rate as smaller size particles. The resulting change of the rel-
ative size distribution is a result of this study, and does not
need to be anticipated in the initial conditions.
Note that, since we do not consider particle growth and
destruction in this study, nor the back reaction of the par-
ticles on the gas, we are free to renormalise the the actual
size distribution of particles a posteriori by changing the
interpretation of the statistical weight factors of the macro-
particles. We can thus match any given initial actual size
distribution, and derive what the final distribution would
be. We also note that since the span in space and time of
our models is very limited (relative to disk extents and life-
times), ignoring coagulation and shattering is an excellent
approximation; such effects may instead be taken into ac-
count by modifying -- a posteriori as per the remarks above
-- the initial particle distribution.
3.1 Equations of motion
= aG + ad,
The acceleration of macro-particles,
dup
(18)
dt
is governed by aG, the differential forces originating from
the planet, the star and the Coriolis force, which, in analogy
with Eq. 15, are
aG = − GMp(xx + yy + zz)
r3
+ 3Ω2
Kxx − Ω2
Kzz − 2ΩKz × u, (19)
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
7
(a) Initial conditions at time = 0.
(b) Relaxed simulation at time = 30 000
Figure 3. Relaxation of initial conditions for the m095t00 run. The panels show density slices and velocity streamlines at the midplane,
with almost the entire extent of the simulation box in the x-direction and 1/5 of the way in the y-direction. Horseshoe orbits are clearly
visible in panel (b). Panel (b) is plotted with a highly enhanced contrast (the colour scale spans only 20% in density) to show the wakes
and density increase at the orbital radius in the y-direction. The white and red circles (barely visible in these figures) denote the Hill
radius and the isothermal Bondi radius of the embryo, respectively. The embryo itself is smaller than a pixel.
(a) Initial conditions at time = 0.
(b) Relaxed simulation at time = 30 000
Figure 4. Relaxation of initial conditions for the m095t00 run. Intermediate scale density slices at the midplane. The large red circle is
the Bondi radius. The solid green circle in panel (a) shows the current 'active' radius, while subsequent dashed circles show where future
'releases' will place a temporary embryo surface. The tiny black circle shows the actual radius of the embryo, which is reached after the
final release.
and the acceleration from the gas drag,
ad =
ugas − upart
ts
,
(20)
which is a product of the relative velocity of the particle,
upart, with respect to that of the gas, ugas, and the inverse of
the stopping time, ts, which, in the Epstein regime is:
ρ•s
ρvth
,
ts =
where ρ• is the solid density (taken to be 3 g cm−3) and vth
(21)
is the mean thermal velocity of the gas. We further assume
that particles are spherical.
For simplicity and speed, we parametrise the stopping
MNRAS 000, 1 -- 21 (2018)
time as:
(cid:18) ts
(cid:19)
(cid:16)
=
s
1cm
(cid:17)(cid:18) cd
(cid:19)
,
(22)
ρ
1 yr
where cd = 10−12 g cm−3 is a normalisation constant, de-
fined such that in the sub-sonic Epstein regime, the stopping
time is 1 year for a 1 cm particle in a gas density of 10−12
g cm−3. This estimates the stopping time in the disk and
outer parts of the Hill sphere sufficiently well, while ignor-
ing the additional decrease of the stopping time due to the
higher temperature near the surface of the embryo. Parti-
cles that reach such depths of the potential well are in any
case doomed to accrete, so nothing is lost by ignoring the
8
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
with solution:
∆t/ts
∆v = −
(1 + ∆t/ts) (v − vg(t + ∆t)) .
(25)
In the limit ∆t/ts (cid:29) 1, the velocity ends up very close to
the gas velocity (as it should), but without the need to take
very short time steps, resulting in a much more efficient time
integration. In the limit ∆t/ts (cid:28) 1, one recovers the normal
kick-drift-kick behaviour.
We thus advance the velocity in the first (kick) step
with:
v1 = v0 +
∆t
(1 + ∆t/ts) (aG(r0, t0) + ad(r0 + v0∆t, t0 + ∆t)) .
(26)
Figure 5. Physical structure of the primordial atmosphere for
run m095t00 at t = 0 (initial conditions; IC) and t = 30 000 =
6.1 years (fully relaxed). The horizontal and vertical cuts overlay
each other nearly perfectly.
temperature dependence, and doing so reduces the cost to
compute the stopping time.
Although this approximation may overestimate the
stopping time slightly, the precise value is not critical, since
there is in any case an uncertain factor that depends on the
shape and porosity of the particles, and a different value can
be compensated for by a corresponding shift of the assumed
particle distribution.
In a similar vein, there is no need to expand the pa-
rameter space with different assumed disk densities, since
that too corresponds to a rescaling of the particle stopping
time, and a corresponding rescaling of the disk and adiabatic
atmospheric structures with a constant factor in density.
Because of its dependence on gas density and particle
size, the stopping time can vary over many orders of mag-
nitude. Given the time step, ∆t, that is necessary based on
the velocity and acceleration of gravity, aG, one can thus
have both ∆t (cid:28) ts and ∆t (cid:29) ts. The latter case means that
the differential equation governing the particle path becomes
stiff.
To handle this, particle positions and velocities are up-
dated with a modified kick-drift-kick leapfrog scheme (e.g.
Dehnen & Read 2011), reformulated to give the correct rela-
tive speed also in the asymptotic limit when the time step is
much larger than the stopping time. Instead of sub-cycling,
which is computationally expensive, we use the method that
is commonly referred to as "forward time differencing", which
ensures that the solution has the correct asymptotic be-
haviour, both when the stopping time is very long and very
short, and that the behaviour is also nearly the correct one
in intermediate cases.
With forward time differencing, the velocity update, as-
suming for now only an acceleration due to the drag force,
ad, may be written:
v(t + ∆t) = v(t) + ∆v = v(t) + ∆t ad(t + ∆t) ,
where the acceleration is time-centred at t +∆t. The resulting
update relation is:
∆v = ∆t ad(t + ∆t) = −(∆t/ts)(v + ∆v − vg(t + ∆t)) ,
(23)
(24)
The particle position and time are then updated (drift step)
with:
r2 = r0 + v12∆t ,
and
t2 = t0 + 2∆t .
Finally, the third step (kick) advances the velocity using the
acceleration at the end of the full time step, with
(28)
(27)
v2 = v1 +
∆t
(1 + ∆t/ts) [aG(r2, t2) + ad(t2)] .
(29)
We have tested this formulation on particle drift in disks
with known solutions (Weidenschilling 1977), recovering be-
haviour sufficiently accurate for our purposes for arbitrary
values of Stokes numbers St = Ωts.
The particle update procedure is computationally ef-
ficient, with the average update cost per cell on an Intel
Ivy Bridge core increasing from 0.65 µs without particles to
0.72 µs with one particle per cell, indicating an average up-
date cost of about 70 nanoseconds per particle. Most of that
time is spend looking up properties needed to calculate the
particle-gas drag force.
3.2 Injection of pebbles
Once the simulations are considered to be fully relaxed, peb-
bles are injected into the system. The probability that a
particular cell in a particular patch will spawn a macro-
particle is proportional to Nppc, the number of particles per
cell (which need not be an integer). We set Nppc to 0.05 in
most runs, but increase it to 1.0 in a few runs to improve
the statistics when studying the distribution of particles in
small volumes.
The mass carried by a macro-particle is encoded in a
'weight factor', w = ρ∆V/Nppc, where ∆V is the cell volume.
The weight factor is thus proportional to the mass of gas
in the cell where the macro-particle originates divided by
Nppc. An advantage of this approach is that the sum over
all particles (or a subset defined by particle size or initial
particle location) results in the total mass of those particles
(to be renormalized by a preferred initial dust-to-gas ratio).
However, because the mass contained in any given cell
can vary substantially, particles may need to be split to
avoid the situation where the particle mass in a patch is
represented by too few particles. For purposes of efficiency,
the need for splitting is only checked when a particle passes
from one patch to another. If an inbound particle carries
more weight than, e.g., five times the initial average in the
MNRAS 000, 1 -- 21 (2018)
ρ (g cm-3)10−1010−910−810−7T (K)102103104r (RP)100101102100101102Density IC horizontal cut vertical cutTemperature IC horizontal cut vertical cutPebble dynamics and accretion onto rocky planets.
9
(a) Midplane.
(b) z = 10 RP.
(c) z = 25 RP.
(d) z = 75 RP.
(e) Vertical component of the vorticity in the
midplane.
(f) xz slice of uz at y = 0.
Figure 6. Gas flows around the M = 0.95 M⊕ embryo in the m095t00 run. Panels (a-d) show the velocity magnitude at different heights
above the midplane. The scale of the colour bar is deliberately held constant. Panel (e) shows the vertical component of the vorticity in
the midplane and panel (f) shows uz in a xz slice at y = 0. Velocity streamlines are also shown.
patch, it is split into a number of "child particles". The child
particles continue to carry, also after any additional splits,
a memory of their initial identity, which is needed for iden-
tifying sub-populations based on initial physical location.
When a macro-particle is created, it is assigned a ran-
dom position inside a cell and a randomly chosen size:
s = 10−3R cm,
(30)
where R ⊆ [0,1] is a random number. The spawned macro-
particles are initially given the gas velocity of the cell. We
spawn ≈ 2.6 million macro-particles in most runs, with 53
million in a few runs with Nppc = 1. We save a complete set of
particle data when a snapshot of the gas dynamics is taken,
but, for visualization purposes, we also separately save tra-
jectory information for a smaller subset of macro-particles
(∼15 000) at every particle time step. These 'tracer parti-
cles' are used to accurately track the trajectories of accreting
particles and correspondingly to show the escape routes of
particles that enter the Hill sphere and then leave again.
4 RESULTS: GAS FLOWS
We begin by investigating the gas flow patterns around the
different mass cores. In the last several years, a number of
authors have studied the hydrodynamics in the vicinity of
embedded planetary embryos (e.g. D'Angelo & Bodenheimer
2013; Ormel et al. 2015b; Fung et al. 2015; Masset & Ben´ıtez-
Llambay 2016; Cimerman et al. 2017; Lambrechts & Lega
MNRAS 000, 1 -- 21 (2018)
2017; Xu et al. 2017, etc.) across a broad range of different
conditions (e.g. different core masses, orbital distances from
the parent star, disk structures), physical processes (isother-
mal/adiabatic/realistic EOS, with/without radiative energy
transfer, magnetic fields, etc.). As illustrated below, our sim-
ulations are consistent with these previous studies, while of-
fering both a higher resolution and a larger domain around
the embryo.
Fig. 6 shows the details of the gas flow in the m095t00
run for the M = 0.95 M⊕ planetary embryo. For clarity, we
zoom-in to the Hill radius (234 embryo radii, white circle).
The red circle denotes the nominal Bondi radius, RB, which
is ≈ 50RP. The embryo is marked by a black circle, but is too
small to be visible in these figures. The first four panels show
the velocity magnitude in the disk midplane (z=0), and at
z = 10, 25, and 75 RP. As expected, the perturbation of the
gas due to the embryo decreases with increasing height above
the midplane, as indicated by the streamlines3. The velocity
magnitude in the horizontal plane 75 planet radii above the
midplane is nearly uniform along the y-direction, indicating
that, at that height, there are no substantial perturbations
where the embryo is located, consistent with Ormel et al.
(2015b). We also observe that the width of the horseshoe
orbits decrease with increasing height above the midplane,
3 Note, however, that the streamlines does not give an entirely
accurate impression of the flow pattern because they only consider
the velocity in the plane of the slice.
10
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
(a) Midplane.
(b) z = 10 RP.
(c) z = 25 RP.
(d) z = 75 RP.
(e) Vertical component of the vorticity in the
midplane.
(f) xz slice of uz at y = 0.
Figure 7. Same as Fig. 6, but for the M=0.096 M⊕ embryo in the m01t00 run.
(a) m01t00 run.
(b) m05t00 run.
(c) m095t00 run.
Figure 8. Velocity magnitude in the midplane in the vicinity of the embryos as a function of embryo mass.
which agrees with the results of Fung et al. (2015), but con-
tradicts the isothermal results of Masset & Ben´ıtez-Llambay
(2016). This might be attributed to the fact that we use a
non-isothermal EOS, but further analysis of the horseshoe
orbits is beyond the scope of this work. Panel (e) of Fig.
6 shows the vertical component of the vorticity measured
relative to the rotating coordinate system. A clear vorticity
pattern can be seen, with separators splitting the flow into
regions that pass the embryo on the left and right sides. The
flows passing on the "wrong" side contribute directly to pos-
itive (prograde, counter-clockwise) vorticity, which adds to
the rotation of the flow.
Panel (f) of Fig. 6 shows the vertical velocity, vz , in
an xz plane at y = 0. A vertical gas flow pattern can be
distinguished: gas flows towards the embryo from the poles
and moves away at the midplane. Zonal flow patterns are
visible as closed streamline loops. Vertical velocities inside
RB are of the order of a few cm s−1, which agrees well with
previous works(e.g. Ormel et al. 2015b; Lambrechts & Lega
2017).
Qualitatively, the M = 0.5 M⊕ case is very similar, thus
the figure panels for this mass are presented in Appendix A;
Fig. A1.
The gas flows around the M = 0.096 M⊕ embryo
(m01t00 run) behave somewhat differently, as illustrated in
Fig. 7. Because of the lower embryo mass, the perturbations
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
11
away from pure shear flow are smaller, and horseshoe orbits
therefore reach further in towards the embryo surface. The
strong features in vorticity are correspondingly smaller. The
vertical flows of the gas, meanwhile, approach the embryo
over a broader extent radially, relative to the size of the Hill
sphere. In our simulations, the gas flow patterns in the in-
nermost region change with time, alternating between domi-
nating gas inflows from the leading/outer horseshoe flow and
the trailing/inner one. The locations of the stagnation points
also vary with time. However, it is possible that part of this
behaviour is related to the recursive refinement of the grid.
We performed exploratory runs with the same mesh refine-
ment, but without the planet embryo, and found root-mean-
square velocities in the innermost region on the order of 10
m s−1. This is similar to the case with the low-mass embryo
present, and hence the small scale, low amplitude motions
are unreliable. These motions are not caused by noise gen-
erated directly by the mesh refinement, but rather by slight
differences in the cancellation of the force of gravity and
pressure gradients, where small differences due to numerical
resolution result in imperfect hydrostatic balance.
4.1 Gas dynamics close to the planetary embryos
One reason why this work stands out relative to previous
studies is that we are able to resolve the detailed gas dynam-
ics (and hence accurate particle paths) in the close vicinity
of the planetary embryos. Fig. 8 shows the gas velocity mag-
nitude in the vicinity of RB for the M = 0.096 M⊕ (panel
(a)), M = 0.5 M⊕ (panel (b)) and M = 0.95 M⊕ (panel
(c)) embryo masses. The M = 0.096 M⊕ core is not massive
enough to form a hydrostatic atmosphere -- the horseshoe
flows constantly flush the atmosphere and only a tiny part
of it (≈ 2 RP from the embryo) is nearly static. With increas-
ing embryo mass, more gas becomes bound and the veloc-
ity magnitude decreases. Nevertheless, there are always slow
gas movements inside RB due to both gas from the disk that
reaches deeper interior layers and from small scale turbu-
lence. Therefore, the primordial atmospheres are never fully
isolated and always interact with the surrounding disk.
5 RESULTS: PARTICLES
An obvious advantage of our large simulation box is the
large particle reservoir. Smaller simulation boxes, especially
with non-periodic x-boundaries, often experience a depletion
of macro-particles. This makes the accurate evaluation of
accretion rates more difficult, and additional injections of
particles may be needed (e.g. Xu et al. 2017).
In Fig. 9, we show time-series of particle mass density
from run m095t10 across a range of integrated radial layers
(shells) from the embryo surface to twice the Hill radius.
Black denotes the first snapshot after the injection of the
pebbles (after 40 time units) while the other lines are plot-
ted at 4 000 time unit intervals (which corresponds to ≈0.8
of an orbital period). We also plot the integrated density of
the gas divided by 100 (the canonical gas-to-dust ratio), in
cyan for the same radial bins. The bottom panel shows the
number of macro-particles in each radial bin. Here, the ini-
tial dips on the black curves stem from changes in the level of
refinement. As time goes on, the number of macro-particles
MNRAS 000, 1 -- 21 (2018)
Figure 9. Mass distribution of solids in radial shells close to
the embryo at different times for the m095t10 run. The vertical
dashed and dotted lines show the Bondi and Hill radii, respec-
tively. The cyan line denotes the gas density in each shell, divided
by a canonical dust-to-gas ratio of 100. The lower panel shows the
number of macro-particles as a function of radius. The bin width
dr is 4.66RP. As initially local macro-particles with low weights
are replaced by macro-particles with larger weights from larger
distances, the statistical noise increases.
Figure 10. Vertical distribution of pebble sizes in the m095t00
run 1.6 orbits after the injection of macro-particles.
decreases due to accretion and transport, but new particles
are continuously coming from further out. These new par-
ticles originate in patches with lower numerical resolution
(i.e. larger cell sizes), and thus carry larger representative
weights. The increasing noise level, meanwhile, is due to the
increasing contribution from these higher statistical weight
macro-particles.
Even after 2.4 orbits, Fig. 9 shows no indication of de-
pletion of solids -- the mass influx through a given shell is
sufficient to replace any accreted or lost particles. This im-
plies that our initial injection of particles is sufficiently nu-
merous to allow accurate measurements of accretion rates
without further injections. Similar trends are seen for the
other embryo masses; as the solid density is proportional to
the gas density, we choose to not show separate figures for
them.
Npart100101102103104105R (RP)100200300400Solid density (g cm-3)10−1310−1210−1110−10100200300400Time after pebble injection (Ω-1)0.0160.821.622.43 gas / 10012
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
(a) Particle distribution over a 2RH vertical
extent above/below the midplane.
(b) Same as (a), but with a 2RB vertical
extent
(c) A zoom-in of panel (b).
Figure 11. Gas density and spatial distribution of particles close to the midplane in the m095t00 run. The particles here are split into
3 size bins: 0.001 < s ≤ 0.01 cm (yellow), 0.01 < s ≤ 0.1 cm (red), and 0.1 < s ≤ 1 cm (green). The black circle denotes the Hill sphere,
while the red circle denotes the Bondi radius.
5.1 The distribution of particles
We initialize particles uniformly in the dust-to-gas ratio, and
we therefore assume the initial vertical stratification of par-
ticles is the same as that of the gas. Particles then settle to-
wards the midplane due to the vertical force of gravity, with
settling time scales of the order of 1/(tsΩ2
time scale for small and intermediate size particles particles
is long relative to the duration of our experiments, while it is
comparable to our run times for the largest particles. Lack-
ing proper turbulence in the disk, the settling of particles is
anyway not very realistic -- turbulence stirs up the particles
and controls their scale height. Additionally, zonal gas flows
(e.g. panel (f) of Fig. 6) close to the Hill sphere are capable
of stirring up well-coupled particles.
K). The settling
Therefore, instead of working with ill-defined and un-
known equilibrium populations, we choose instead well-
defined,
localized sub-populations with initially constant
dust-to-gas ratios for our measurements of accretion rates.
Tests have shown that it takes only ∼0.5 orbits after the in-
jection of particles for the measured accretion rates to stabi-
lize. The hydrodynamic flow patterns are established quickly
as well, taking approximately the same amount of time. This
shows that our initial relaxation of the flow patterns, which
was allowed to continue for ∼6 orbital periods, was certainly
sufficient and allows for accurate measurements of accretion
rates.
Fig. 10 shows the vertical distribution of pebbles in the
m095t00 run, 1.6 orbits after the injection of macro-particles.
The smallest particles (s ≤ 0.01 cm; dark colours) continue
to show essentially the vertical distribution of the gas, while
larger particles (s > 0.1 cm; green to yellow) have begun to
settle. This figure also nicely illustrates how small a pertur-
bation the Hill sphere is relative to the size of our computa-
tional domain.
Fig. 11 shows the spatial distribution of particles and
their sizes at the midplane, two orbits after the injection of
particles. This time, the particles are split into 3 size bins:
s ≤ 0.01 cm (yellow), 0.01 ≤ s < 0.1 (red) and s > 0.1 cm
(green). The vertical extent of particles we display is limited
to 1RH in panel (a) and 2RB in panels (b,c), above and be-
low the midplane. One instantly notices the two leading and
trailing arms, stretching away from the embryo. The large
concentration of particles there indicates that these particles
are moving away from the embryo very slowly. The wakes
are located just inside the separatrix sheets of the horse-
shoe orbit, with the smallest particles mostly concentrating
in the outermost regions and the larger ones lying closer
to the orbit of the embryo. These macro-particles carry a
relatively small statistical weight, because they originate at
patches with high resolution and small mass per cell, so the
features are somewhat artificially enhanced in the Figure,
but it therefore also nicely illustrates the preferred 'leaving'
paths of the particles. A large fraction of these particles en-
ter but then later leave the Hill sphere, and the features are
therefore concentrated in the midplane.
A zoom-in to RB (panel (c)) reveals that there is a
concentration of small particles (s < 0.1 cm) inside the
Bondi sphere. Larger particles do not linger there for long
since they are effectively accreted. The figure shows particle
paths based on high-cadence positions of tracer-particles,
sub-divided into 3 size bins. The green colour represents
0.001 ≤ s < 0.01 cm, blue 0.01 ≤ s < 0.1 cm and black
s0.1 ≤ s < 1 cm size particles. The smallest particles spend
around 0.1Ω−1 inside the Bondi sphere before they reach
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
13
(a) m095t10
(b) m05t10
(c) m01t10
Figure 12. Trajectories of accreted particles. The axes are given in units of embryo radii. Green, blue and black paths are for particle
sizes 0.001 ≤ s < 0.01 cm, 0.01 ≤ s < 0.1 cm, and 0.1 ≤ s < 1 cm, respectively.
the embryo. In contrast, particles with s ≥ 0.01 cm, are ac-
creted on average 10 times faster, after they reach the Bondi
sphere. Alternatively, some of the smallest particles, if they
linger far enough from the embryo, are later carried out by
the horseshoe gas flows.
In these runs we do not consider convection driven by
the accretion heating (but see below), particle growth and
evaporation. Sufficiently strong convection could possibly
prevent the smallest particles from reaching the core. Also,
as the gas temperature close to the embryo (≈ 4500 K near
the M = 0.95 M⊕ embryo) greatly exceeds the melting point
of any solid, evaporation of solids and the resulting enrich-
ment of the atmosphere with heavy elements (e.g. Alibert
2017; Brouwers et al. 2017), is bound to play an important
role in models with a realistic treatment of evaporation. As
shown by Fig. 1, a more realistic equation of state leads to
lower temperatures at the base of the atmospheres, but still
large enough to cause evaporation.
Fig. 13 similarly shows the spatial distribution of peb-
bles close to the midplane (inside 2RB) in the m01t00 run.
It shows a very similar picture -- escaping particles are
trapped in the leading/inner and trailing/outer horseshoe
orbit parts. The zoom-in of the Figure (panel b) reveals an
interesting, but expected difference -- there is no accumu-
lation of particles inside the Bondi sphere. There are two
factors that contribute to this: the gas flows may be strong
enough to readily carry small and intermediate size parti-
cles away, and larger particles are rapidly accreted. Panel
(c) in Fig. 12 reveals the answer: the stopping time, even of
the smallest size particles we consider in our runs, is short
enough that the gravitation of the embryo is not damped
by the gas drag enough to prohibit particles from being ac-
creted. The smallest particles in our simulation take merely
a 0.04 Ω−1 to be accreted. And the number of these small-
est particles crossing the Bondi sphere is much larger, com-
pared to the M = 0.95 M⊕ embryo. Here we remind the
reader that we do not trace all the particles with this high
cadence, therefore only a fraction of all the accreted parti-
cles are shown in Fig. 12 (we also, for greater clarity, limit
the number of traces shown in these figures to 150).
Fig. 12 shows evidence for a variation of hydrodynam-
MNRAS 000, 1 -- 21 (2018)
ical deflection (e.g. Sellentin et al. 2013) - higher mass
embryos are accompanied by much denser primordial at-
mospheres, in which the stopping times become smaller.
Whereas the majority of the particles in the distribution
we consider here have nearly ballistic orbits when being ac-
creted by the lowest mass embryo (Fig. 12, panel (c)), they
become significantly affected by the higher gas density en-
velopes in panels (b) and (a). In the latter, the highest-mass
embryo case, the motion of the gas almost completely con-
trols the motion of particles with size s < 0.01 cm.
5.2 Accretion of pebbles
When pebbles approach an embryo they are subject to an
increasing two-body force and start to veer towards the em-
bryo. Although from Fig. 12 it might seem that they come
from all over the place, the figure greatly under-represents
the particles (150 traces selected out of 15 000, again ran-
domly selected from ≈2.6 million macro-particles), and the
scales are concentrated to the closest 10-20 planetary radii
around the embryos. The full ensemble of trace-particles
in fact show distinct inflows and outflows through the Hill
sphere.
Fig. 14 plots the mass accretion rates across the Hill
sphere, Bondi sphere and the embryo surface over the course
of the m095t10 run, displayed as a set of Hammer projec-
tions. The macro-particles are sub-divided into 3 represen-
tative size-bins (0.0 ≤ s ≤ 0.01 cm, 0.01 < s ≤ 0.1 cm and
0.1 < s ≤ 1.0 cm). The figures show that the bulk of the
net accretion comes from near the midplane, and predom-
inantly from the large size particles. As they do not feel
much gas drag, some of the largest particles can and do
cross the Hill sphere from arbitrary directions. The domi-
nating inflow regions are associated with the inner/trailing
and outer/leading horseshoe orbits, and corresponding out-
flows in the outer/trailing and inner/leading horseshoe or-
bits. The accretion rates at distances RB are much more
isotropic, without clearly preferred trajectories. The modest
outflows are mainly close to the midplane region. Not all
of the mass accreted through the Bondi sphere reaches the
embryo surface during the time scales we consider here. The
14
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
the figures for M= 0.5 M⊕ and M = 0.096 M⊕, embryos are
in Appendix A.
To better illustrate the dependence of accretion on par-
ticle size, an animation showing the trajectories of pebbles
in the vicinity of an M = 0.95 M⊕ embryo (m095t10) is avail-
able on-line as supplementary material (and on YouTube 4).
In the animation, cyan represents 0.0 ≤ s ≤ 0.01 cm parti-
cles, red represents 0.01 < s ≤ 0.1 cm particles and yellow
represents 0.1 < s ≤ 1.0 cm particles. Velocity streamlines
are shown in green. The coloured circles meanwhile denote
the Hill radius (magenta) and the "nominal" effective accre-
tion radii for unperturbed flow, reff (Lambrechts & Johansen
2012, eq. 40; see below) for the three, aforementioned size
bins. The animation clearly illustrates that: (i) particles that
pass within their respective reff are accreted; (ii) small, well-
coupled particles are particularly affected by the gas flow
and often do not reach reff.
5.3 Quantitative measurements of accretion rates
To allow selection of arbitrary sub-populations, our initial
particle population is distributed all over space, with weights
proportional to the local cell gas mass. To obtain accretion
rates pertaining to a disk where particles have settled to-
wards the midplane, we select the sub-population that ini-
tially resides within one Hill radius from the midplane, and
subsequently measure accretion rates based on only those
particles, renormalising their macro-particle weights w so
they correspond to a total mass equal to the assumed dust-
to-gas ratio (0.01) times the total gas mass in the experi-
ment.
By sub-dividing the particle distribution into 6 size bins,
we can study the dependency of the accretion rates on size.
Fig. 15 shows that the net accretion rate onto the embryo
scales linearly with the size s, which is different than the s2/3
scaling predicted by e.g. Ormel & Klahr (2010); Lambrechts
& Johansen (2012), assuming unperturbed shear flow condi-
tions. As acknowledged in Lambrechts & Johansen (2012),
these conditions only hold when the mass of the planet em-
bryo is very small. The key reason this scaling does not hold
in our simulations is because the flow pattern (i.e. horseshoe
orbits with stagnation points/separatrix surfaces) and den-
sity (with contrasts of the order of 103) in the vicinity of
the planet are very different from unperturbed shear flow.
Small, well-coupled particles are most strongly affected by
these differences; they tend to follow gas streamlines, and
therefore follow horseshoe orbits or are deflected around the
planet. As a result, the dependence of the accretion rate on
particle size becomes steeper than 2/3. The measured linear
scaling is valid even for larger particles, all the way to dm-
sized particles, as illustrated in Fig. 16. The accretion rate
saturates (no longer increases with particle size) for particles
larger than about 1 meter.
A crude measure for the importance of flow effects for
pebble accretion can be obtained by equating the pebble ac-
cretion impact radius in the shear limit (reff ∼ τ1/3RH, where
τ is the Stokes number) with the Bondi radius RB -- the scale
on which flow effects are expected to become important. For
reff < RB flow effects may be expected to change accretion
4 https://youtu.be/2GNgoc71qIw
MNRAS 000, 1 -- 21 (2018)
(a) Spatial distribution of particles over a
2RB vertical extent above/below the mid-
plane. The black circle denotes the Hill
sphere.
(b) A zoom-in of panel (a) to the Bondi
sphere. The red circle denotes the Bondi
sphere.
Figure 13. Gas density and spatial distribution of particles close
to the midplane in the m01t00 run. The pebbles here are split into
3 size bins: yellow s ≤ 0.01 cm, red 0.01 < s ≤ 0.1cm and s > 0.1
cm are green.
missing mass is contained in the smallest particles, which
trace the gas flows very closely. Having even the slightest
turbulence in the primordial atmosphere (which is always
the case, as discussed earlier) traps these particles and pro-
hibits them from reaching the surface over sufficiently small
time scales, but nonetheless, they eventually are accreted.
That said, as the experiments approach a steady-state, the
net accretion rate across a sphere of a given radius (e.g. RH
or RB) centred on the embryo approaches parity with the
accretion rate onto the embryo. The two top rows of Fig. 14
also show how inefficiently the subsets of smaller particles
are accreted. Very similar trends are present in the other
two embryo mass cases we consider in this work, therefore,
Pebble dynamics and accretion onto rocky planets.
15
Figure 14. From left to right, the mass accretion rate across the Hill sphere, Bondi sphere and embryo surface, respectively, over the
course of the m095t10 run, displayed as a Hammer projection. From top to bottom are the particle accretion rates for three size bins
0.0 ≤ s ≤ 0.01 cm, 0.01 < s ≤ 0.1 cm and 0.1 < s ≤ 1.0 cm. A longitude of 270◦ corresponds to the -y-direction and a latitude of 0◦
corresponds to the disc midplane.
Figure 15. Pebble accretion rates onto the embryo, for parti-
cle sizes within each size bin, drawn from an initial distribution
consisting of particles within ±RH from the midplane, for the
three embryo masses, 0.95 M⊕(black), 0.5 M⊕(red), and 0.096
M⊕(blue). The green line shows a mass accretion rate propor-
tional to particle size, s.
rates that have been derived under the assumption of an
unperturbed gas flow. Equating the radii, we find a thresh-
old mass of MP∗ ∼ h3τ1/2Mstar, where h is the disk aspect
ratio. Hence above MP∗ flow effects may modify pebble ac-
cretion estimates based on assumptions of pure shear flow.
This expression also states that smaller particles (smaller
MNRAS 000, 1 -- 21 (2018)
Figure 16. Pebble and boulder accretion rates onto the embryo,
for particle sizes between 0.1 - 100 cm, subdivided into 6 size bins,
drawn from an initial distribution consisting of particles within
±RH of the midplane for the 0.95 M⊕ embryo mass case.
τ) are more severely affected, qualitatively consistent with
what we find. However, to explain the reduction quantita-
tively, i.e., to explain the linear dependence of (cid:219)M on τ that
we find, requires a more thorough analysis, which will be
considered in a future work.
The runs with 56 million macro-particles give similar
results as the corresponding runs with 2.6 million particles,
with less than 10% difference in the accretion rates through
Accretion rate [M⊕ yr-1]10−810−710−610−5Particle size [cm]10−310−210−1100M = 0.95 M⊕M = 0.5 M⊕M = 0.096 M⊕Accretion rate [M⊕ yr-1]10−510−4Particle size [cm]10−110010110216
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
100%, except for particle sizes below 0.003 cm, which are still
very well coupled to the gas there. Over size bins where the
efficiencies are statistically well determined, they also scale
approximately linearly with size. This is a consequence of
the denominator (the unsigned, radially inward transport
rate) being approximately independent of particle size (at a
given radius) in our regime.
We do not see significant systematic differences when
different pressure bump parameters ζp are used for the same
mass embryo. This is to be expected, since the results will
only change when the radial settling has become significant.
One would expect on the one hand a tendency for reduction
of the accretion rates, because of the smaller y-velocities
closer to the orbital radius of the embryo, and on the other
hand a tendency for increase, due to the accumulation of
large size particles. Ultimately, the accretion rates would
be constrained by the radial influx of particles from larger
orbital radii into the pressure trap.
6 ACCRETION DRIVEN CONVECTION
The high resolution close to the embryo surfaces enables us
to resolve convection driven by accretion heating. Accretion
of solids onto the embryo releases a lot of potential energy,
which is converted to heat via the friction force. To simulate
the effects of the heating due to the accretion, we added an
extra heating term Qaccr to one of our simulations, m095t10-
conv. Considering a shell of extent dr at some distance r
from the embryo, the volume dV = Adr, where A = 4πr2,
then has a potential energy difference (which is per unit
mass) given by Φ(r + dr) − Φ(r). To get an energy per unit
volume we need to multiply with the mass accreted per unit
time:
(cid:219)MGMp
4πr4 ,
Qaccr =
where Qaccr is the heating rate per unit volume and (cid:219)M is the
mass accretion rate (which we set to 10−6 M⊕ yr−1). The
heating rate, Qaccr, is then added to the entropy equation
(Eq. 4) via:
(31)
∂s
∂t
= ... +
ρQaccr
Pgas
.
(32)
Fig. 18 illustrates the consequences of the addition of the
accretion heating term in the m095t10-conv run. It shows
extensive convective patterns, which considerably modify
the gas flows around the embryo (left panel). The horse-
shoe orbits have somewhat receded, and the weak random
flows that appear in non-convective runs are effectively over-
whelmed by the much stronger convective flow patterns. The
middle panel shows a vertical projection of the convective
motions, which extend out to nearly 40 embryo radii. The
vicinity of the embryo is no longer in a quasi steady state -
the gas flow patterns constantly change, and the intercon-
nection between the primordial atmosphere and the disk is
thus strengthened; i.e., no parts of the atmosphere are effec-
tively isolated. The convection is very dynamic, with local
velocities reaching ∼1 km s−1, inducing sound waves when
the convective cells reach the outskirts of the atmosphere.
The direction of the outflows are constantly changing and
now lack any significant constant patterns, but do have an
MNRAS 000, 1 -- 21 (2018)
Figure 17. Local pebble accretion efficiency (the net accretion
rate divided by the radially inward mass flux across a sphere
of a particular radius) as a function of particle size, for spheres
of several different radii, drawn from an initial distribution of
particles within ±RH of the midplane for the m095t10 run.
the Hill sphere and less than 1% difference at the Bondi
sphere. Thus the results are rather robust, even when us-
ing only 2.6 million macro-particles. Naturally, however, the
greater the number of particles, the smaller a fraction of the
population one can sample from and still retain statistical
significance; i.e., one has the ability to perform more focused
studies of particle behaviour, as a function of size and initial
position. The M = 0.95 M⊕ case, for example, initially has
only about 20% of the macro-particles within one Hill radius
of the midplane. Thus, to study midplane accretion rates it
is useful to have a large particle population available.
An estimate of the maximum accretion rate may be
obtained by taking the dust-to-gas ratio times the average
gas mass flux through the Hill sphere. For our M = 0.95
M⊕ embryo immersed in our 1/10 MMSN disk, this flux is
≈ 4 10−3 M⊕yr−1. With a dust-to-gas ratio of 0.01, we thus
estimate a maximum accretion rate of about 4 10−5 M⊕yr−1.
This rate is exceeded by about a factor of five in the 0.3 -- 1 m
size bin in Fig. 16. By comparison, the 'Hill rate', ΣpΩKR2
H,
is 3.9 10−3 M⊕yr−1 and nearly identical to the actual, 3-D
gas mass flux.
The accretion rates for the largest size bin (∼0.3 -- 1
cm) are approximately 3 10−6 M⊕ yr−1, 7 10−6 M⊕ yr−1 and
2 10−5 M⊕ yr−1, for the three embryo masses, respectively
(cf. Fig. 15). The accretion rates scale, as expected, approxi-
mately as the Hill rate time the Stokes number (proportional
to M2/3a−3/4 with MMSN scalings).
Fig. 17 illustrates the local efficiency of accretion of par-
ticles across particular spheres in the m095t10 simulation.
The 'local accretion efficiency' is defined as the net accre-
tion rate through a sphere of radius R, normalised by the ac-
cretion rate obtained by counting only the solids transport
inwards through the same sphere (note that this definition
differs from other uses of "accretion efficiency" in the liter-
ature; cf. Lambrechts & Johansen 2012; Guillot et al. 2014;
Ormel 2017). Evidently, a large fraction of all the particles
entering the Hill sphere leave again without being accreted.
As illustrated in Fig. 17, the local efficiency is particularly
low (near 0%) for the lowest mass particles, which are well
coupled to the gas and, due to temporal fluctuations in the
accretion rate, can in fact temporarily show negative values.
The local efficiency, meanwhile, rapidly grows for smaller
radial distances, and at ∼30RP the local efficiency is nearly
Accretion efficiency00.20.40.60.81Particle size [cm]10−310−210−1100RH100 RPRB35 RP30 RPPebble dynamics and accretion onto rocky planets.
17
Figure 18. From left to right: xy slice in the midplane, yz slice at x = 0 of the entropy per unit mass and particle paths in the vicinity of
the core in a single snapshot of accretion driven convection in the m095t10-conv run. For clarity, in the middle panel, the gas streamlines
are omitted. In the right panel, black curves show traces for particles of size 0.1 < s ≤ 1 cm, blue 0.01 < s ≤ 0.1 and green 0.001 < s ≤ 0.01
cm, respectively.
approximately isotropic distribution of the velocity disper-
sion.
When we see or model convection at a stellar surface
(e.g. in the form of 'granulation' on the Sun), the cellular
patterns that develop are very much constrained by conser-
vation of mass. In a stratification where mass density drops
very rapidly with height, mass conservation forces the flows
to "turn over", from up to down, which again forces the hor-
izontal size to be limited (lest the ratio of horizontal to ver-
tical speed become very large). Here, we have a situation
that is quite different, with a "small" body at the centre
of a large spherical volume. This situation allows the flows
much greater freedom, with a possibility of expansion in all
directions.
The heating driving the convection in this study, com-
puted assuming (cid:219)M = 10−6 M⊕yr−1, is not consistent with the
actual, about 20 times larger measured accretion rates (e.g.
Fig. 15). Moreover, radiative energy transfer and a more
realistic EOS would tend to change the properties of con-
vection, so to reach true consistency, improvements are nec-
essary. Nevertheless, it is possible to learn already from the
current experiment, which can illustrate to what extent the
convective motions are able to affect accretion rates relative
to an analogous non-convective case.
The right hand panel of Fig. 18 shows a small subset of
high cadence paths for accreted particles, colour coded by
size. Although their paths are affected, the largest particles
are not significantly stirred up by the convective motions.
The smaller the particles are, however, the more they are
affected by the convective motions. Quantitative measure-
ments show that the average of the unsigned mass fluxes
through spherical surfaces where convection is active is much
larger than in the non-convective case.
However, as illustrated in Fig. 19, we find that convec-
tion in the current case does not significantly reduce the net
accretion rate onto the embryo. Formally, the local accretion
efficiency (the net mass accretion rate as a fraction of the
unsigned, radially inward solid mass flux; Fig. 20) is reduced
significantly for smaller particles. This is because convective
motions gives rise to larger unsigned rates of radial mass
transport, and hence increase the denominator in the defi-
MNRAS 000, 1 -- 21 (2018)
Figure 19. Pebble accretion rates in the accretion driven convec-
tion run (m095t10-conv). Particle sizes are split into 6 representa-
tive logarithmic size bins. The green line shows a mass accretion
rate proportional to particle size, s.
Figure 20. Local pebble accretion efficiency in the accretion
driven convection run (m095t10-conv). Particle sizes are split into
6 representative logarithmic size bins.
nition of local accretion efficiency. The convective motions
also cause larger temporal fluctuations, and as a result we
measure negative net accretion rates and a local accretion
efficiency of less than zero at RB for some of the smallest size
bins. The convective velocities in this case are generally on
the order of the free fall drift velocities of the particles, as
illustrated in Fig. 21, but the systematic drift of particles rel-
Accretion rate [M⊕ yr-1]10−810−710−610−5Particle size [cm]10−310−210−1100RH100 RP30 RP10 RP Ṁ ∝ sAccretion efficiency00.20.40.60.81Particle size [cm]10−310−210−1100RH100 RP30 RP10 RP18
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
while feedback from the particle ensemble on the gas was
neglected -- consistent with the modest dust-to-gas ratios oc-
curring in the models. The motions of a subset of particles
were recorded with high cadence, allowing for their trajec-
tories and histories to be studied in great detail.
Our main results may be summarized as follows:
(i) Starting from approximately hydrostatic initial condi-
tions, relaxation periods on the order of one orbit are suf-
ficient to establish near quasi-stationary conditions in the
vicinity of the Hill sphere.
(ii) The relaxed atmospheric structures do not differ
dramatically from spherically symmetric hydrostatic atmo-
spheres matched to surrounding disk conditions at the Hill
radius.
(iii) Adiabatic and hydrostatic atmospheres with an ideal
gas and γ = 1.4 do not differ substantially from atmospheres
computed with a more realistic, tabular EOS (Tomida &
Hori 2016).
(iv) The presence of convection changes conditions well
inside the Hill sphere from near-steady flows to dynamically
evolving flows, albeit stationary statistics (e.g. velocity dis-
persion as a function of radius) develop in less than an or-
bital period.
(v) The convective motions result in an enhanced mass
exchange between different layers inside the Hill sphere, but
do not significantly affect the mass exchange through the
Hill sphere with the surroundings, nor do they affect the net
accretion rates significantly.
(vi) In our fiducial 1/10 MMSN model, the (unsigned)
gas mass flux through the Hill sphere for our 3D horseshoe
flows is about 4 10−3 M⊕ yr−1, and thus agrees closely with
H = 3.9 10−3 M⊕ yr−1 (obtained
the canonical "Hill rate", ΣΩR2
assuming Σ = 170 g cm−2 and RH/R⊕ = 231).
(vii) The gas mass flux, scaled with the dust-to-gas ratio,
is reflected in similar rates of particles passing both in and
out of the Hill sphere, with inflow increasingly dominate for
increasing particle size.
(viii) As illustrated by Figs. 13 and 14, the particles that
are not accreted and leave the Hill sphere have preferred
paths along the separators of the horseshoe flow and are
concentrated towards the midplane.
(ix) In contrast, at the canonical Bondi sphere, the parti-
cle fluxes are mainly inward (except for the smallest particle
sizes), and hence changes in the detailed structure of the
layers close to the embryo (e.g. due to the EOS, including
from evaporation of pebbles) are not important for the final
accretion rates.
(x) The accretion rates for particles is observed to scale
linearly with particle size, different than for unperturbed
shear flow conditions, where it scales as s2/3 (Ormel & Klahr
2010; Lambrechts & Johansen 2012, 2014). We attribute the
steepening of the power-law to aerodynamic deflection of
well-coupled particles under the higher densities and very
different flow patterns in our simulations relative to unper-
turbed shear flow conditions.
(xi) We find that the accretion rate of 0.3 -- 1 cm parti-
cles for an 0.95 M⊕ embryo immersed in a 1/10 MMSN disk,
assuming a dust-to-gas ratio of 0.01 and assuming that the
dust has settled to a midplane layer with height H < RH,
is about 2 10−5 M⊕yr−1. Since the accretion rates scale es-
sentially linearly with particle size and inversely with gas
MNRAS 000, 1 -- 21 (2018)
Figure 21. Free fall drag velocities of particles of size 1 cm (blue),
0.1 cm (green), 0.01 cm (red) and 0.001 cm (cyan). The velocity
dispersion of the m095t10-conv run is shown in magenta.
ative to the gas continues, and since the convective motions
do not result in a net transport of mass, their effect on the
particle drift is a second-order effect. It is possible, however,
that this could change if the convective motions were suf-
ficiently resolved and included more turbulent, small-scale
motions.
The effect on the particle accretion rate is also likely
to increase, and possibly become quite significant, when ac-
cretion heating is included self-consistently by including ra-
diative energy transfer and a realistic EOS. The resulting,
self-consistent accretion rates are likely to be lower than the
current values, but the convective motions are likely to be
stronger than in our current, nominal case.
7 SUMMARY AND CONCLUSIONS
In this paper, we have reported the results from high resolu-
tion, nested grid, 3D hydrodynamic and particle simulations
which focus on solid accretion onto Mars- to Earth-mass
planetary embryos. The primary goals of this work were to
establish realistic three-dimensional flow patterns inside and
outside the Hill sphere, to use these to determine accurate
accretion rates, and to investigate how the accretion rates
depend on embryo mass, pebble size, and other factors.
Our excellent spatial resolution and efficient time in-
tegration, using locally determined time steps, made it
possible -- using only a single, 20-core compute node per
run -- to perform the first-ever simulations of pebble accre-
tion covering scales from several disk scale heights to a few
percent of a planet radius, including atmospheres that self-
consistently connect to a surrounding accretion disk. It is
also the first time that accretion heating driven convection
has been modelled in a domain with such a large dynamic
range.
We conducted a series of simulations with different plan-
etary embryo masses and strengths of a pressure trap. To
be able to accurately measure accretion rates, and to study
how they are affected by convection, we injected millions
of macro-particles with representative sizes ranging from
10 micron to 1 m. The motion of the particles were fol-
lowed, taking into account drag forces relative to the gas,
Pebble dynamics and accretion onto rocky planets.
19
tening of the hydrostatic structure inside the Hill sphere
could lead to the formation of a circumplanetary disk with
prograde rotation (cf. Johansen & Lacerda 2010). In order
to model such an evolution, it is crucial to realistically in-
clude the major sources of heating, energy transport, and
cooling: Accretion heating, the resulting convection, and ra-
diative cooling while also using a realistic EOS (e.g. Tomida
et al. 2013; cf. Fig. 1) and opacities.
To this end, we are currently conducting simulations
that include these effects. In the future, we will also model
the accretion heating in a manner consistent with the solid
accretion rate (cf. Sect. 6) and include pebble destruction
via ablation (e.g. Brouwers et al. 2017; Alibert 2017).
ACKNOWLEDGEMENTS
We would like to thank the anonymous referee for in-depth
and helpful comments that have helped to significantly im-
prove the quality of this manuscript. The work of AP and
AN was supported by grant 1323-00199B from the Dan-
ish Council for Independent Research (DFF). CWO is sup-
ported by the Netherlands Organization for Scientific Re-
search (NWO; VIDI project 639.042.422). The Centre for
Star and Planet Formation is funded by the Danish National
Research Foundation (DNRF97). Storage and computing re-
sources at the University of Copenhagen HPC centre, funded
in part by Villum Fonden (VKR023406), were used to carry
out the simulations presented here. The authors are grateful
to Kengo Tomida and Yasunori Hori, who kindly provided
the tabular EOS. AP is grateful to the Anton Pannekoek
Institute at the University of Amsterdam for a three month
stay during which part of this work was carried out. This
research made use of the YT project for analysis and visu-
alisation (Turk et al. 2011).
density, the accretion rate for the same size particles would
be essentially the same in higher density disks; the larger
supply of solids in a denser disk will tend to be cancelled by
a correspondingly slower accretion speed.
(xii) For our 1/10 MMSN solar mass disk, the accretion
rate saturates (no longer increases with particles size) for
particles larger than about 1 m, while for denser disks the
particles can be even larger. Thus, only for such large par-
ticle sizes will the accretion rates onto an embryo increase
with the disk surface density. Conversely, if mm-size pebbles
dominate the mass budget, the accretion rates depend only
weakly on the disk surface density.
Having such a high spatial resolution means that we
could only cover a limited parameter space and run the sim-
ulations for a relatively small number of orbital periods. We
have also relied on an ideal gas EOS, while in reality the
high temperatures near the embryo surface would lead to
the dissociation of molecular hydrogen (e.g. D'Angelo & Bo-
denheimer 2013), and therefore result in a much denser and
somewhat cooler inner atmosphere. This would not signifi-
cantly alter the accretion efficiency of the larger particles,
however, as these effects occur at small radii from which
these particles are not able to escape.
Given the accretion rates determined in Section 5.3, we
can estimate the growth time scales from low mass seeds to
full planets, under the specified conditions (i.e. a gas surface
density of 170 g cm−3, a dust-to-gas ratio of 0.01, and solids
from within H < RH. As per the results discussed above,
to lowest order, the effects of disk density on mass supply
and accretion speed cancel. The remaining dominant scaling
is the dependence on embryo mass, which is approximately
M2/3, corresponding to a mass that grows approximately
as t3. This implies that, as long as this approximate scaling
persists, growth from a small seed takes on the order of three
times the instantaneous mass divided by the instantaneous
accretion rate.
We conclude that if growth from low mass embryos to
full size planets is dominated by accretion of 0.3 -- 1 cm size
particles, the accretion from a low mass seed would have
taken of the order of 0.15 million years in the case of Earth,
and of the order of 0.1 million years in the case of Mars.
With chondrule size particles -- if we take them to be 0.3-1
mm -- the growth times are 10 times longer. These estimates
assume a local ratio of dust-to-gas surface densities of 0.01.
Increased dust-to-gas ratios, e.g., due to radial accumulation
of solids in a pressure trap, would lead to correspondingly
reduced accretion times.
During the growth phase, embryos would inevitably be
surrounded by hot, primordial atmospheres of much larger
mass than the current day atmospheres of the rocky Solar
System planets. The decrease in disk densities required to
stop pebble accretion would also result in removal of much
of the primordial atmosphere as it expanded to match the
decreasing pressure at its outer boundary (Nordlund 2011;
Schlichting et al. 2015; Ginzburg et al. 2016). The isotopic
signatures present in noble gases in the Earth's atmosphere
may provide indirect evidence related to this (Pepin 1991,
1992a,b).
A decrease in gas and dust densities will lead to in-
creased rates of radiative cooling, at first dominated by cool-
ing in the vertical direction. The consequential, gradual flat-
MNRAS 000, 1 -- 21 (2018)
20
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
REFERENCES
ALMA Partnership et al., 2015, ApJ, 808, L3
Alibert Y., 2017, Astronomy and Astrophysics, 606, A69
Andrews S. M., et al., 2016, ApJ, 820, L40
Bell C. P. M., Naylor T., Mayne N. J., Jeffries R. D., Littlefair
S. P., 2013, MNRAS, 434, 806
Bitsch B., Johansen A., Lambrechts M., Morbidelli A., 2015a,
A&A, 575, A28
Bitsch B., Lambrechts M., Johansen A., 2015b, A&A, 582, A112
Bizzarro M., Connelly J. N., Krot A. N., 2017, in Pessah M.,
Gressel O., eds, Astrophysics and Space Science Library
Vol. 445, Astrophysics and Space Science Library. p. 161,
doi:10.1007/978-3-319-60609-5 6
Bollard J., et al., 2017, Science Advances, 3, e1700407
Brouwers M. G., Vazan A., Ormel C. W., 2017, preprint, 1708,
arXiv:1708.05392
Casassus S., et al., 2013, Nature, 493, 191
Cassan A., et al., 2012, Nature, 481, 167
Chatterjee S., Tan J. C., 2014, ApJ, 780, 53
Cimerman N. P., Kuiper R., Ormel C. W., 2017, MNRAS, 471,
4662
D'Angelo G., Bodenheimer P., 2013, ApJ, 778, 77
Dehnen W., Read J. I., 2011, European Physical Journal Plus,
126, 55
Ebisuzaki T., Imaeda Y., 2017, New Astron., 54, 7
Friedrich J. M., Weisberg M. K., Ebel D. S., Biltz A. E., Corbett
B. M., Iotzov I. V., Khan W. S., Wolman M. D., 2015, Chemie
der Erde / Geochemistry, 75, 419
Fung J., Artymowicz P., Wu Y., 2015, ApJ, 811, 101
Ginzburg S., Schlichting H. E., Sari R., 2016, ApJ, 825, 29
Greenberg R., Hartmann W. K., Chapman C. R., Wacker J. F.,
1978, Icarus, 35, 1
Guillot T., Ida S., Ormel C. W., 2014, A&A, 572, A72
Hayashi C., 1981, in Sugimoto D., Lamb D. Q., Schramm D. N.,
eds, IAU Symposium Vol. 93, Fundamental Problems in the
Theory of Stellar Evolution. pp 113 -- 126
Hubickyj O., Bodenheimer P., Lissauer J. J., 2005, Icarus, 179,
415
Ida S., Makino J., 1993, Icarus, 106, 210
Inaba S., Ikoma M., 2003, A&A, 410, 711
Isella A., P´erez L. M., Carpenter J. M., Ricci L., Andrews S.,
Rosenfeld K., 2013, ApJ, 775, 30
Johansen A., Lacerda P., 2010, MNRAS, 404, 475
Johansen A., Mac Low M.-M., Lacerda P., Bizzarro M., 2015,
Science Advances, 1, 1500109
Kokubo E., Ida S., 1998, Icarus, 131, 171
Kritsuk A. G., et al., 2011, ApJ, 737, 13
Lambrechts M., Johansen A., 2012, A&A, 544, A32
Lambrechts M., Johansen A., 2014, A&A, 572, A107
Lambrechts M., Lega E., 2017, Astronomy and Astrophysics, 606,
A146
Levison H. F., Thommes E., Duncan M. J., 2010, AJ, 139, 1297
Lommen D., Maddison S. T., Wright C. M., van Dishoeck E. F.,
Wilner D. J., Bourke T. L., 2009, A&A, 495, 869
Masset F. S., Ben´ıtez-Llambay P., 2016, ApJ, 817, 19
Meru F., Juh´asz A., Ilee J. D., Clarke C. J., Rosotti G. P., Booth
R. A., 2017, ApJ, 839, L24
Morbidelli A., Nesvorny D., 2012, A&A, 546, A18
Nordlund A., 2011, in Sozzetti A., Lattanzi M. G., Boss A. P.,
eds, IAU Symposium Vol. 276, The Astrophysics of Planetary
Systems: Formation, Structure, and Dynamical Evolution. pp
105 -- 112, doi:10.1017/S1743921311020023
Nordlund A., Galsgaard K., Stein R. F., 1994, in Rutten R. J.,
Schrijver C. J., eds, NATO Advanced Science Institutes (ASI)
Series C Vol. 433, NATO Advanced Science Institutes (ASI)
Series C. p. 471
Nordlund A., Ramsey J. P., Popovas A., Kuffmeier M., 2018,
preprint, (arXiv:1705.10774)
Ormel C. W., 2017, in Pessah M., Gressel O., eds, Astrophysics
and Space Science Library Vol. 445, Astrophysics and Space
Science Library. p. 197, doi:10.1007/978-3-319-60609-5 7
Ormel C. W., Klahr H. H., 2010, A&A, 520, A43
Ormel C. W., Kuiper R., Shi J.-M., 2015a, MNRAS, 446, 1026
Ormel C. W., Shi J.-M., Kuiper R., 2015b, MNRAS, 447, 3512
Ormel C. W., Liu B., Schoonenberg D., 2017, A&A, 604, A1
Paardekooper S.-J., Mellema G., 2006, A&A, 453, 1129
Pepin R. O., 1991, Icarus, 92, 2
Pepin R. O., 1992a, Annual Review of Earth and Planetary Sci-
ences, 20, 389
Pepin R. O., 1992b, in Luhmann J. G., Jakosky B. M., eds, LPI
Contributions Vol. 787, Evolution of the Martian Atmosphere.
Piso A.-M. A., Youdin A. N., Murray-Clay R. A., 2015, ApJ, 800,
82
Pollack J. B., Hubickyj O., Bodenheimer P., Lissauer J. J.,
Podolak M., Greenzweig Y., 1996, Icarus, 124, 62
Rein H., Lesur G., Leinhardt Z. M., 2010, A&A, 511, A69
Rubanenko L., Steinberg E., Schlichting H., Paige D. A., 2017, in
AAS/Division for Planetary Sciences Meeting Abstracts #49.
p. 413.01
Schlichting H. E., Sari R., Yalinewich A., 2015, Icarus, 247, 81
Sellentin E., Ramsey J. P., Windmark F., Dullemond C. P., 2013,
A&A, 560, A96
Stevenson D. J., 1982, Planet. Space Sci., 30, 755
Tan J. C., Chatterjee S., Hu X., Zhu Z., Mohanty S., 2016, IAU
Focus Meeting, 29, 6
Testi L., Natta A., Shepherd D. S., Wilner D. J., 2003, A&A, 403,
323
Tomida K., Hori Y., 2016, personal communication
Tomida K., Tomisaka K., Matsumoto T., Hori Y., Okuzumi S.,
Machida M. N., Saigo K., 2013, ApJ, 763, 6
Turk M. J., Smith B. D., Oishi J. S., Skory S., Skillman S. W.,
Abel T., Norman M. L., 2011, The Astrophysical Journal Sup-
plement Series, 192, 9
Visser R. G., Ormel C. W., 2016, A&A, 586, A66
Weidenschilling S. J., 1977, MNRAS, 180, 57
Whipple F. L., 1972, in Elvius A., ed., From Plasma to Planet.
p. 211
Xu Z., Bai X.-N., Murray-Clay R. A., 2017, The Astrophysical
Journal, 847, 52
Zeng L., Sasselov D. D., Jacobsen S. B., 2016, ApJ, 819, 127
de Val-Borro M., et al., 2006, MNRAS, 370, 529
van der Marel N., et al., 2013, Science, 340, 1199
APPENDIX A: ADDITIONAL FIGURES
Fig. A1 shows the details of the gas flow in the m05t00 run
for the M = 0.5 M⊕ planetary embryo. The similarity with
the higher mass case is apparently because embryos of such
masses have well established primordial atmospheres, which
are very close to being hydrostatic, leading to very similar
flows in the vicinity of the Hill sphere.
Fig. A2 shows the mass accretion rates over the course
of the m05t10 run. The accretion rates are about a factor
of two lower than in the m095t10 case, and the preferred
accretion paths through the Hill sphere are once again closer
to the midplane and the stagnation points (separators).
Fig. A3 shows the mass accretion rates over the course
of the m01t10 run. Note that the contour levels are modi-
fied, in order to enhance the features. The simulation shows
a similar scenario to the cases with higher mass embryos, ex-
cept the mass accretion rates are once again smaller. Also,
there is effectively no loss of mass from RB. This is because
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
21
(a) Midplane.
(b) z = 10 RP.
(c) z = 25 RP.
(d) z = 75 RP.
(e) Vertical component of the vorticity in the
midplane.
(f) xz slice of the vertical velocity component
at y = 0.
Figure A1. Same as in Fig. 6, but for the M=0.5 M⊕ embryo from the m05t00 run.
the RB for such a small mass embryo is just ∼12.3 embryo
radii, or just 4% of RH, and therefore actually deeper in the
potential well of the embryo than in the higher mass cases.
This paper has been typeset from a TEX/LATEX file prepared by
the author.
MNRAS 000, 1 -- 21 (2018)
22
A. Popovas, A. Nordlund, J. P. Ramsey, C. W. Ormel
Figure A2. Same as Fig. 14, but for the m05t10 run.
MNRAS 000, 1 -- 21 (2018)
Pebble dynamics and accretion onto rocky planets.
23
Figure A3. Same as Fig. 14, but for the m01t10 run.
MNRAS 000, 1 -- 21 (2018)
|
1707.03345 | 1 | 1707 | 2017-07-11T16:15:53 | The GTC exoplanet transit spectroscopy survey. VII. An optical transmission spectrum of WASP-48b | [
"astro-ph.EP"
] | We obtained long-slit optical spectroscopy of one transit of WASP-48b with the Optical System for Imaging and low-Intermediate-Resolution Integrated Spectroscopy (OSIRIS) spectrograph at the 10.4 m Gran Telescopio Canarias (GTC). We integrated the spectrum of WASP-48 and one reference star in several channels with different wavelength ranges, creating numerous color light curves of the transit. We fit analytic transit curves to the data taking into account the systematic effects present in the time series in an effort to measure the change of the planet-to-star radius ratio ($R_p/R_s$) across wavelength. After removing the transit model and systematic trends to the curves we reached precisions between 261 ppm and 455-755 ppm for the white and spectroscopic light curves, respectively. We obtained $R_p/R_s$ uncertainty values between $0.8 \times 10^{-3}$ and $1.5\times 10^{-3}$ for all the curves analyzed in this work. The measured transit depth for the curves made by integrating the wavelength range between 530 nm and 905 nm is in agreement with previous studies. We report a relatively flat transmission spectrum for WASP-48b with no statistical significant detection of atmospheric species, although the theoretical models that fit the data more closely include of TiO and VO. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. GTC_Wasp48
August 29, 2018
c(cid:13)ESO 2018
7
1
0
2
l
u
J
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
4
3
3
0
.
7
0
7
1
:
v
i
X
r
a
The GTC exoplanet transit spectroscopy survey VII
An optical transmission spectrum of WASP-48b
F. Murgas1,2,3,4, E. Pallé1,2, H. Parviainen1,2, G. Chen1,2,7, L. Nortmann1,2, G. Nowak1,2, A. Cabrera-Lavers1,5, and
N. Iro6
1 Instituto de Astrofísica de Canarias (IAC), E-38205 La Laguna, Tenerife, Spain
2 Departamento de Astrofísica, Universidad de La Laguna (ULL), E-38206 La Laguna, Tenerife, Spain
3 Univ. Grenoble Alpes, IPAG, F-38000 Grenoble, France
4 CNRS, IPAG, F-38000 Grenoble, France
5 Gran Telescopio Canarias (GTC), E-38712, Breña Baja, La Palma, Spain
6 Theoretical Meteorology group, Klimacampus, University of Hamburg, Grindelberg 5, 20144 Hamburg, Germany
7 Key Laboratory of Planetary Sciences, Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China
Received 04 April 2017/ Accepted 22 June 2017
ABSTRACT
Context. Transiting planets offer an excellent opportunity for characterizing the atmospheres of extrasolar planets under very different
conditions from those found in our solar system.
Aims. We are currently carrying out a ground-based survey to obtain the transmission spectra of several extrasolar planets using the
10 m Gran Telescopio Canarias. In this paper we investigate the extrasolar planet WASP-48b, a hot Jupiter orbiting around an F-type
star with a period of 2.14 days.
Methods. We obtained long-slit optical spectroscopy of one transit of WASP-48b with the Optical System for Imaging and low-
Intermediate-Resolution Integrated Spectroscopy (OSIRIS) spectrograph. We integrated the spectrum of WASP-48 and one reference
star in several channels with different wavelength ranges, creating numerous color light curves of the transit. We fit analytic transit
curves to the data taking into account the systematic effects present in the time series in an effort to measure the change of the planet-
to-star radius ratio (Rp/Rs) across wavelength. The change in transit depth can be compared with atmosphere models to infer the
presence of particular atomic or molecular compounds in the atmosphere of WASP-48b.
Results. After removing the transit model and systematic trends to the curves we reached precisions between 261 ppm and 455-755
ppm for the white and spectroscopic light curves, respectively. We obtained Rp/Rs uncertainty values between 0.8×10−3 and 1.5×10−3
for all the curves analyzed in this work. The measured transit depth for the curves made by integrating the wavelength range between
530 nm and 905 nm is in agreement with previous studies. We report a relatively flat transmission spectrum for WASP-48b with no
statistical significant detection of atmospheric species, although the theoretical models that fit the data more closely include of TiO
and VO.
Key words. planetary systems -- techniques: spectroscopy -- planets and satellites: atmospheres
1. Introduction
With more than 2700 confirmed transiting extrasolar planets1
and many unconfirmed candidates, the characterization of their
atmospheres through transmission spectroscopy has become an
active subfield in exoplanetary studies. Although the first de-
tection of elements using transmission spectroscopy were made
from space (e.g., Charbonneau et al. 2002, Vidal-Madjar et al.
2004, Lecavelier Des Etangs et al. 2008, Désert et al. 2009),
ground-based telescopes and instruments are also able to achieve
the high precision measurements required by this technique.
Indeed, after the first space-based discoveries of atoms and
molecules in extrasolar planets, many ground-based observa-
tions followed (e.g., Snellen et al. 2008, Redfield et al. 2008,
Bean et al. 2011, Sing et al. 2012, Bean et al. 2013) and at present
there are several efforts to continue the characterization of exo-
planetary atmospheres from the ground.
We are currently carrying out a transmission spectroscopy
survey to characterize the atmospheres of transiting planets using
1 http://exoplanet.eu/catalog/
long-slit spectroscopy with the 10.4 m telescope Gran Telesco-
pio Canarias (GTC). The principle of transmission spectroscopy
is to measure the change in the planetary radius at different wave-
lengths during a transit event, and to use this change to infer the
presence of particular components of the exoplanet atmosphere.
Because of their relatively big atmospheric scale heights, hot
Jupiters are excellent targets for atmospheric detection from the
ground and they are the main targets of our survey (e.g., Murgas
et al. 2014, Parviainen et al. 2016, Pallé et al. 2016, Nortmann
et al. 2016, Chen et al. 2017).
The hot Jupiter WASP-48b was discovered by Enoch et al.
(2011) as part of the Wide Angle Search for Planets (WASP,
Pollacco et al. 2006). Its host star is a slightly evolved F star
with an apparent magnitude of V = 11.72 ± 0.14 mag, a mass
of M = 1.19 ± 0.05 M(cid:12), a stellar radius of R = 1.75 ± 0.09 R(cid:12),
and an effective temperature of Te f f = 6000 ± 150 K. Enoch
et al. (2011) measured a planetary mass of M = 0.98 ± 0.09
MJup, a radius of R = 1.67 ± 0.10 RJup, and an orbital period
of 2.14 days. Sada et al. (2012) measured a planet-to-star ra-
dius ratio in the J band of Rp/Rs = 0.0988+0.0051
−0.0049, which agrees
Article number, page 1 of 8
A&A proofs: manuscript no. GTC_Wasp48
with the value reported in the discovery paper by Enoch et al.
(2011) (Rp/Rs = 0.0980 ± 0.001). O'Rourke et al. (2014) used
Spitzer to measure the secondary transit of WASP-48b in the H,
Ks, 3.6 µm, and 4.5 µm bands. For the dayside emission spectra
they fitted a blackbody model with an effective temperature of
Te f f = 2158 ± 100 K. Using the models of Fortney et al. (2008)
and Burrows et al. (2008) to fit their data, they deduced that
WASP-48b has a weak or absent temperature inversion. Ciceri
et al. (2015) presents a follow up of this system using multicolor
broadband photometry; this work updated the stellar parameters
of WASP-48 (M = 1.062 ± 0.074 M(cid:12), R = 1.519 ± 0.007 R(cid:12)),
measured a smaller planetary radius (R = 1.396 ± 0.051 RJup),
and found a flat transmission spectrum using broadband filters.
Turner et al. (2016) presents ground-based near-UV data of sev-
eral transiting systems including WASP-48b, and found a Rp/Rs
for the U band of 0.0916± 0.0017, a shallower transit depth than
the one found by Enoch et al. (2011).
This paper is organized as follows. In §2 and §3 we describe
the observations and data reduction, in §4 we describe the light
curve fitting process, and in §5 we show the transmission spec-
trum of WASP-48b. Finally, in §6 we present a discussion of the
results and in §7 the conclusions of this paper.
2. Observations
The GTC is a 10.4 m telescope located at Observatorio Roque de
los Muchachos in La Palma, Spain. For the observations, the Op-
tical System for Imaging and low-Resolution-Integrated Spec-
troscopy (OSIRIS, Cepa et al. 2000) in its long-slit spectroscopic
mode was chosen.
The data were taken on July 17, 2014; the binning was set to
2 × 2; the readout speed of OSIRIS was chosen to be 200 KHz
with a gain of 0.95 e−/ADU; and readout noise of 4.5 e−. The
observations were made using the R1000R grism (spectral cov-
erage of λ ∼ 520 − 1040 nm) and with a custom built slit with a
width of 40 arcsec. The use of a wide slit minimizes the system-
atic effects that may appear due to flux losses caused by seeing
variations during the observations. The exposure time was set to
15 seconds, thus assuring a high signal-to-noise ratio below the
saturation level of the detector. Close to 50 minutes of data be-
fore and after the transit were collected; the total time of data
acquisition was 5 hours, which translated into 552 science im-
ages. During the observations, the airmass varied from 1.32 to
1.22 (minimum airmass of 1.19, maximum telescope elevation
angle was 63.2◦).
During the transit of WASP-48b, OSIRIS was able to obtain
spectra for the target and four reference stars (Figure 1) simul-
taneously. Because of its low brightness, three of the stars were
not suitable to be used as reference and were discarded from the
analysis. All the results presented here were produced using the
flux of the star R1 as reference (see Table 1).
The universal time of data acquisition was determined us-
ing the recorded headers of the images. The opening and closing
time of the shutter was used to compute the time of mid expo-
sure. Then, using the code written by Eastman et al. (2010)2,
the mid exposure time was converted to the Barycentric Julian
Date in Barycentric Dynamical Time (BJD_TDB). All the re-
sults present here were produced using BJD as the time standard.
2 http://astroutils.astronomy.ohio-state.edu/time/
Article number, page 2 of 8
Fig. 1. GTC OSIRIS image through the slit (top) and calibrated science
image (bottom). In this image, the target is the star labeled T and the
selected reference star is labeled R1. Other reference stars in the field
(R2, R3, and R4) are marked, but were not used in the data analysis due
to their faint magnitude.
3. Data reduction
The raw data were calibrated using standard procedures. With
the calibration images provided by the GTC team, we proceeded
to create an average bias using 32 images. We requested a large
number of flats in order to have a good estimation of the pixel
sensitivy of OSIRIS; the flat field calibration set was composed
of 100 images that were combined after subtracting the average
bias. The master flat was then normalized by fitting (including
rejection of deviant pixels) a high degree Chebyshev polynomial
to the flux across the dispersion axis.
Following Murgas et al. (2014), the extraction and wave-
length calibration of the spectra was done using a customized
PyRAF script that calls different IRAF3 tasks that work with
long-slit spectroscopic data. A fixed aperture size of 20 pixels
in width (5.08 arcsec) was used to extract the spectra. This aper-
ture width was chosen after determining which one delivered the
lowest scatter in the out-of-transit points in the white light curve
(wavelength range integrated between 530-905 nm). In order to
correct for drifts along the dispersion axis, the cross-correlation
between the first image and the rest of the time series for each
star was computed with IRAF's task FXCOR. This process was
repeated between the target and the reference star. Additionally,
a Gaussian function was fitted to the spectral profile of each star
in the time series to monitor drifts perpendicular to the disper-
sion axis. The computed full width at half maximum (FWHM) of
the spatial profile for each star was used in the fitting procedure
as a proxy of the seeing variations observed during the transit.
An example of the extracted spectrum and wavelength coverage
of GTC R1000R grism can be seen in Figure 2.
3 IRAF is distributed by the National Optical Astronomy Observatory,
which is operated by the Association of Universities for Research in
Astronomy (AURA) under a cooperative agreement with the National
Science Foundation.
CCD 10100200300400500600700800Pixel0500100015002000PixelTR1R2R3R4F. Murgas et al.: The GTC exoplanet transit spectroscopy survey VII
Table 1. Observed stars.
Star
2MASS name
WASP-48
R1
J19243895+5528233
J19243079+5527485
RA (FK5)
DEC (FK5)
19h 24m 38.984s +55◦ 28m 23.39s
19h 24m 30.787s +55◦ 27m 48.52s
B (mag) V (mag) K (mag)
10.37
12.31
11.99
10.45
11.65
11.44
905 nm), and 6 curves made with a bin size of 10 nm (covering
the spectral region 564.5 nm - 624.5 nm).
Figure 3 presents the flux of WASP-48, the reference star R1,
and the flux ratio between the target and R1 for the white light
curve. The plot shows clearly that variations in the flux of the tar-
get and reference star were present during the observations, but
they are mostly corrected by the differential photometry. How-
ever, there are some residual systematic effects still present in
the light curve which need to be taken into account in the fitting
procedure in order to have a robust error estimation.
4.1. Model selection
As in Nortmann et al. (2016), the Bayesian Information Criterion
(BIC) was used to select the best light curve model. The BIC
penalizes a high χ2 value and/or the number of free parameters
used in the fitting procedure; the model that presents the lowest
BIC is the one selected.
The BIC values for three models were computed for 16 light
curves produced using a bin size of 25 nm of width covering a
wavelength range between 530 nm and 905 nm. The following
models were tested:
M1 = T (p)(a0 + b1t + c0Fw),
M2 = T (p)(a0 + b1t + c0Fw + c1Fw2),
(1)
(2)
M3 = T (p)(a0 + b1t + b2t2 + c0Fw),
(3)
where T (p) is the transit model and p are the transit dependent
parameters, t is the time of the observations, and Fw is the full
width at half maximum (FWHM) of the spectrum profile. All the
tested models present a synthetic transit model with a time de-
pendency to reproduce a second-order color effect not corrected
by the differential photometry, and a FWHM dependent poly-
nomial to model the systematic effects produced by seeing vari-
ations. The FWHM correction was introduced because seeing
variations affected the number of pixels in which the major part
of the stellar flux is contained, and this pixel-dependent effect
will act as a systematic correlated with FWHM (see Nortmann
et al. 2016, Chen et al. 2017). With a slit 40 arcsec in width, the
flux losses produced by not being able to measure the flux con-
tained in the wings of the spectral profile are negligible. Other
possible sources for the systematic effects were tested (e.g., po-
sition of the stars, telescope rotation angle as in Nortmann et al.
2016, etc.), but no strong correlation was found and they were
discarded as parameters for the tested models.
As shown in Figure 4 no model was selected unanimously;
for each curve created there were some bins that preferred a
model (i.e., lowest BIC value) that did not get the majority of
preferences. Since a unique model that presented the lowest BIC
value for all the curves evaluated in this test could not be found,
the function that provided the lowest BIC values at the highest
rate was selected; in this case Model 1. Using a model with a
Article number, page 3 of 8
Fig. 2. Extracted R1000R grism spectrum of WASP-48 (red) and its
reference star (blue). The spectra are not corrected for instrumental re-
sponse or flux calibrated. The shaded gray areas indicate the custom 25
nm passbands used to create the spectroscopic light curves. The broad-
band filters V, R, I, and z − sloan are also plotted in an arbitrary scale
to show the wavelength coverage of the spectrum.
Fig. 3. Individual light curves of the target and reference star (R1). In
addition, the flux ratio between the target and the reference star is plot-
ted for the white light curve. The fluxes are normalized to the first data
point of the time series and with an arbitrary offset in the y-axis.
4. Light curve fitting
To create the light curves, the spectra of the target and reference
star were integrated in different wavelength ranges and with dif-
ferent bin sizes. Then, the integrated flux of the target was di-
vided by the flux of the reference star.
The curves created and analyzed in this work are as follows:
1 white light curve (flux integrated between 530 nm and 905
nm), 15 curves using a bin size of 25 nm (between 530 nm and
5006007008009001000Wavelength (nm)0.00.51.01.52.0Flux (counts ×105)VRIzWASP-48R10.400.450.500.550.600.65BJD - 2456856.0 (days)0.960.970.980.991.001.011.021.031.04Normalized FluxTargetR1T/R1A&A proofs: manuscript no. GTC_Wasp48
Table 2. Transit parameters uniform prior ranges.
Transit Parameter
Uniform prior range
Tc − 2456856 [days]
[0.512256, 0.533096]
Rp/Rs
u1
u2
a/Rs
i [deg]
[0.03, 0.15]
[−1.0, 1.0]
[−1.0, 1.0]
[3.2, 5.8]
[75, 90]
hood of the curves is given by the comparison between the data
and the transit model plus the systematic trends (see Eq. 1), and it
was computed together with red noise estimation using Gaussian
processes with a simple exponential kernel (the George package
for Python, Ambikasaran et al. 2014). The likelihood of the limb
darkening coefficients was computed using PyLDTK and it pro-
vides a more robust way of establishing the coefficient values
instead of the traditional approach of using tables (Parviainen
et al. (2016)). Only uniform priors were used for the parameters
fitted in this work (see Table 2 for the transit parameters priors).
The fitting procedure was split in three stages. First, we per-
formed a global optimization of F = lnL + ln Priors, i.e., the
likelihood weighted by the priors, using PyDE6. The second part
of the process consisted in running a small MCMC (60 inde-
pendent chains, 10000 iterations) as a burn-in period, using the
global optimization results as a seed. Finally, we used the pa-
rameter vector that delivered the highest F value from stage 2 as
a starting point to generate 250 independent chains and ran the
MCMC procedure for 11500 iterations using emcee. After this
process was over, the autocorrelation length in the chains was
calculated in order to avoid strongly correlated parameter val-
ues in the posterior probability distribution. The final values and
uncertainties were obtained by computing the percentiles of the
distributions to get the median and 1σ limits for each parameter.
Following Parviainen et al. (2016), the two red noise param-
eters used to describe the exponential kernel (the amplitude and
time scale), where computed for the white light curve and fixed
for the rest of the curves (i.e., the 25 nm and 10 nm bins).
5. Results
5.1. White light curve
The white light curve and best fitted model are shown in Fig-
ure 5. The peak-to-peak RMS of the residuals (middle panel of
Fig. 5) is 261 ppm. The fitted transit parameters and their cor-
responding 1σ uncertainties are listed in Table 3. The measured
planet-to-star radius ratio agrees with a shallower transit than the
reported values by Enoch et al. (2011) (Rp/Rs = 0.098 ± 0.001)
and Sada et al. (2012) (Rp/Rs = 0.0988+0.0051
−0.0049); we speculate that
the reason for this discrepancy could be due to spots in the stellar
surface.
5.2. Transmission spectrum
The transmission spectrum of WASP-48b was obtained by mea-
suring the change of transit depth across wavelength. Figure 6
shows the time series and best fitted model for the curves made
using an integration bin of 25 nm of width. Table 4 presents
the measured planet-to-star radius ratio and 1σ uncertainties for
each bin. For all the curves analyzed in this work, the uncertain-
ties in the planet-to-star radius ratio range between 0.8×10−3 and
6 https://github.com/hpparvi/pyde
Fig. 4. Frequency of the minimum BIC values for each proposed model.
The model that presents the lowest BIC is the one selected. The BIC
values were computed using curves created by integrating the flux of
the stars using bins of 25 nm in width.
high number of parameters increases the risk of overfitting the
data; however, for most of the bins where the BIC was evalu-
ated the numerical difference between Models 1, 2, and 3 was
not significant.
4.2. Fitting procedure
In order to produce the synthetic light curves, we used the curve
generator code PyTransit4 (Parviainen 2015). This code presents
optimized Python routines that implement the Giménez (2006)
transit models.
For the transit curves, a quadratic limb darkening law was
adopted. The limb darkening coefficients were modeled using
the Python Limb Darkening Tool Kit (PyLDTK5; Parviainen &
Aigrain 2015), this package uses Husser et al. (2013) stellar li-
braries to compute the coefficients using custom passbands. The
stellar parameters used as input for PyLDTK were taken from
the discovery paper of WASP-48b by Enoch et al. (2011): Te f f =
6000 ± 150 K, log g = 4.50 ± 0.15, and [Fe/H] = −0.12 ± 0.12.
For each curve, we set the following transit parameters as
free: the planet-to-star radius ratio Rp/Rs, the quadratic limb
darkening coefficients u1 and u2, the central time of transit Tc,
the orbital semi-major axis over stellar radius a/Rs, and the or-
bital inclination i. The orbital eccentricity and period were fixed
to 0 and 2.14363544 days (Ciceri et al. 2015), respectively. In
addition to the free transit parameters, we also set as free the co-
efficients used to model the systematic trends present in the light
curves.
In order to estimate robustly the values of the modeled pa-
rameters, a Bayesian fitting process was used. The procedure de-
scribed here is similar to the one presented in Parviainen et al.
(2016). A likelihood function was evaluated iteratively using the
Python MCMC suite emcee (Foreman-Mackey et al. 2013). The
likelihood function was given by:
lnL = lnLcurve + lnLLD,
(4)
where Lcurve is the likelihood of the transit light curve and LLD
is the likelihood of the limb darkening coefficients. The likeli-
4 https://github.com/hpparvi/PyTransit
5 https://github.com/hpparvi/ldtk
Article number, page 4 of 8
Model 1Model 2Model 30246810121416FrequencyBIC minimum valueFilters 25nmF. Murgas et al.: The GTC exoplanet transit spectroscopy survey VII
Table 4. Measured Rp/Rs for the 25 nm light curves.
Center (nm)
542.5
567.5
592.5
617.5
642.5
667.5
692.5
717.5
742.5
767.5
792.5
817.5
842.5
867.5
892.5
Rp/Rs
0.09357 ± 0.00128
0.09301 ± 0.00126
0.09301 ± 0.00105
0.09311 ± 0.00117
0.09277 ± 0.00104
0.09292 ± 0.00098
0.09352 ± 0.00103
0.09307 ± 0.00102
0.09301 ± 0.00094
0.09290 ± 0.00102
0.09220 ± 0.00097
0.09249 ± 0.00090
0.09171 ± 0.00097
0.09165 ± 0.00103
0.09082 ± 0.00089
Table 5. Measured Rp/Rs for the 10 nm light curves.
Center (nm)
569.5
579.5
589.5
599.5
609.5
619.5
Rp/Rs
0.09186 ± 0.00144
0.09278 ± 0.00142
0.09249 ± 0.00136
0.09394 ± 0.00122
0.09227 ± 0.00147
0.09378 ± 0.00128
To analyze our transmission spectrum results, we computed
theoretical atmosphere models using Exo-Transmit (Kempton
et al. 2016). With this code, we created three cloud-free mod-
els with solar composition using an isothermal temperature pro-
file and including Rayleigh scattering, but varying the atmo-
spheric composition. In particular, we computed three models:
i) a model atmosphere with TiO and VO, ii) a model atmosphere
with VO (no TiO), and iii) a model atmosphere with no TiO and
no VO.
5.3. Refractory element signatures
The curves made with a bin size of 10 nm were used to put con-
straints on the presence of sodium in the atmosphere of WASP-
48b. This bin size was chosen to cover the Na i 589.0 and 589.6
nm doublet (Murgas et al. 2014).
The bins near the Na doublet (left and right of the bin cen-
tered at the wavelength of the line) were used to compute a
weighted mean Rp/Rs in order to employ them as a continuum
level to see whether there was an extra absorption at the line
core. We calculated a continuum level of
Rp/Rs
= 0.09340 ± 0.00093,
Avg
(cid:16)
(cid:16)
(cid:17)
(cid:17)
(cid:16)
Na
(cid:17)
while the planet-to-star radius ratio at the Na doublet is
Rp/Rs
= 0.09249 ± 0.00136.
The difference between the line and continuum Rp/Rs is
∆
Rp/Rs
= −0.00091 ± 0.03694,
Avg−Na
meaning that we do not have a statistically significant detection
of sodium.
Article number, page 5 of 8
(6)
(7)
(8)
Fig. 5. GTC/OSIRIS WASP-48b white light transit curve. The red
points present the observed time series and the blue points show the
light curve after removing the systematic effects and red noise compo-
nent. The black line represents the best fit determined using our MCMC
analysis. The black dots at the bottom are the residuals of the best fit
with an arbitrary offset.
Table 3. MCMC results of the white light transit curve of WASP-48b.
Tc − 2456856 [days]
Parameter
Rp/Rs
u1
u2
a/Rs
i [deg]
e
Value
0.09330 ± 0.00088
0.43583 ± 0.00183
0.14877 ± 0.00466
0.522666(234)
4.763 ± 0.079
82.474 ± 0.336
0 (Fixed)
1.5× 10−3, while the standard deviations of normalized residuals
vary between 261 ppm for the white light curve, and 455-755
ppm for spectroscopic light curves. The expected photon noise
for the white light curve is 63 ppm and for the spectroscopic
light curves it is in the range of 213-373 ppm, meaning that the
measured noise level of the white light and spectroscopic curves
is close to 4 and 2 times the expected photon noise level, respec-
tively.
Figure 7 presents the transmission spectrum of WASP-48b.
In this figure the gray shadowed region corresponds to ±3 at-
mospheric scale heights (H). One scale height is the distance in
which the atmospheric pressure decreases by a factor e (Euler's
number) and is given by
H =
kBT p
µgp
,
(5)
where kb is the Boltzmann constant, µ is the mean molecular
weight, T p is the temperature of the planet, and gp is the planet's
surface gravity. The atmospheric scale for WASP-48b is H =
614.9 km (using µ = 2.3 times the proton mass, T p = 2000 K,
and gp = 11.5 m/s2 taken from Ciceri et al. 2015).
0.400.450.500.550.600.65BJD - 2456856.0 (days)0.9850.9900.9951.0001.0051.0101.015Relative Fluxσ=261.3 ppmObservedCorrectedA&A proofs: manuscript no. GTC_Wasp48
Fig. 6. Light curves obtained using the filters of 25 nm of width for WASP48-b. Left panel: Observed light curves and best fit (black line). Right
panel: Transit light curves after removing systematic effects and best fit (black line).
At our resolution the potassium doublet (λ 766.48 nm and
769.89 nm) is blended with some atmospheric telluric lines, thus
making a detection of an excess in the transit depth difficult due
to imperfect telluric correction (Parviainen et al. 2016). In this
work none of the curves analyzed here presents a significant ex-
cess in the transit depth for the bins centered in this line. For the
future, the use of a higher resolution grism could help resolve
the blended lines and help improve the telluric correction near
the K doublet.
6. Discussion
Several planetary atmosphere models, supported by observations
of low mass stars, predict the presence of TiO and VO in hot
Jupiters (Seager & Sasselov 1998, Hubeny et al. 2003, Bur-
rows et al. 2007). Moreover, TiO has been proposed as the ab-
sorber responsible for the temperature inversion layer observed
in some transiting gas giants (Fortney et al. 2008). In optical
wavelengths, Désert et al. (2008) reported a detection of TiO
and VO in the atmosphere of HD 209458b, which is supported
by previous albedo measurements for this planet made by Rowe
et al. (2006). Hoeijmakers et al. (2015) searched for the signa-
ture of TiO in HD 209458b using cross-correlation techniques
applied to high resolution spectra, finding no statistically sig-
Article number, page 6 of 8
nificant detection of TiO, although they pointed out that their
technique depends on the accuracy of TiO line lists used to gen-
erate synthetic planet spectrum. A more recently discovered exo-
planet with tentative detection of TiO and VO in the optical is the
highly inflated WASP-127b (Lam et al. 2017, Palle et al. 2017),
although more conclusive studies need to be made. The detection
of TiO and VO molecules in the optical is difficult because their
features are thought to be relatively broad and of low amplitude,
in addition the presence of hazes in the planetary atmospheres
can affect the strength of the features (Pont et al. 2013, Kreid-
berg et al. 2014, Burrows 2014 and references therein).
O'Rourke et al. (2014) measured the secondary transit of
WASP-48b in the infrared (H, Ks, 3.6 µm, and 4.5 µm bands). By
comparing their data to the models of Fortney et al. (2008) and
Burrows et al. (2008), they deduced that WASP-48b has a weak
or absent temperature inversion and moderate day-night energy
circulation. The model that presented the best fit was one with-
out TiO, but models with a more efficient day-night circulation
and including TiO could also provide a reasonably close match
to the data.
Using χ2 statistics to fit the models to the observed transmis-
sion spectrum, we found that the preferred atmosphere model
was the one including TiO and VO (top panel in Fig. 7) with a
reduced chi square of χ2
r = 0.32 (15 degrees of freedom); for
0.400.450.500.550.600.65BJD - 2456856.0 (days)0.880.900.920.940.960.981.001.021.04Relative Flux542.5 nm567.5 nm592.5 nm617.5 nm642.5 nm667.5 nm692.5 nm717.5 nm742.5 nm767.5 nm792.5 nm817.5 nm842.5 nm867.5 nm892.5 nm0.400.450.500.550.600.65BJD - 2456856.0 (days)σ=465.5 ppmσ=461.6 ppmσ=455.7 ppmσ=471.0 ppmσ=461.6 ppmσ=487.7 ppmσ=489.4 ppmσ=516.2 ppmσ=494.4 ppmσ=556.9 ppmσ=559.4 ppmσ=588.6 ppmσ=622.5 ppmσ=678.3 ppmσ=754.7 ppmF. Murgas et al.: The GTC exoplanet transit spectroscopy survey VII
Fig. 7. Transmission spectrum of WASP-48b. The solid lines are atmosphere models computed with Exo-Transmit for a WASP-48b analog with
Teq = 2000 K, isothermal temperature profile, and Rayleigh scattering by H2 in the blue. The dashed gray line represent the Rp/Rs found for the
white light curve and the shaded area is ±3 atmospheric scale heights (H = 614.9 km) above and below this level.
comparison a straight line model gave χ2
r = 0.59. For the full
observed wavelength range, the model including TiO and VO
presented a lower χ2
r value, although it is not statistically sig-
nificant to be differentiated from a featureless spectrum, i.e, a
flat line. The model with only VO and the model without TiO
and VO, had a reduced chi square of χ2
r = 2.27,
respectively.
r = 0.85 and χ2
After 710 nm the large-scale structure of the TiO bands
has a decreasing slope (e.g., Sharp & Burrows 2007, Fortney
et al. 2010), which coincides with the change in slope in the
observed transmission spectrum. We divided the transmission
spectrum in two regions: bins with λ < 710 nm and bins with
λ ≥ 710 nm, and we performed a simple linear fit for both re-
gions. The measured slope in the region with λ < 710 was con-
sistent with a flat line (slope value of (−0.320 ± 8.545) × 10−6),
while the bins with λ ≥ 710 nm presented a fitted slope value
of (−12.57 ± 5.94) × 10−6. Although our data seems to show
two regions with different slopes, the spectral resolution and un-
certainties of the observed transmission spectrum of WASP-48b
does not allow us to claim a detection of TiO and VO.
7. Conclusions
We present here an analysis of a primary transit of WASP-48b
taken with GTC/OSIRIS instrument. Using a time series of long-
slit spectra of WASP-48 and one reference star, we created sev-
eral color light curves in order to explore the change in transit
depth across wavelength.
In this work we analyzed 1 white light curve (wavelength
range 530-905 nm), 15 curves of 25 nm in width (wavelength
range 530-905 nm), and 6 curves of 10 nm (wavelength range
564.5-624.5 nm). All the curves were fitted using a Bayesian
MCMC procedure using a transit model with a quadratic limb
darkening law and reproducing the systematic effects present in
the curves. The uncertainties reported in this work take into ac-
count time correlated noise of unknown origin (red noise) com-
puted using Gaussian processes.
We report a flat, featureless transmission spectrum confirm-
ing previous broadband observations. The obtained transmission
spectrum agrees with the expected cloud-free theoretical atmo-
sphere model that includes the presence of TiO and VO, although
evidence of these molecules in the atmosphere of WASP-48b is
not statistically significant enough to claim a detection. Explor-
Article number, page 7 of 8
0.0900.0920.0940.0960.098Rp/Rs1xsolar TiO VO0.0900.0920.0940.0960.098Rp/Rs1xsolar VO500550600650700750800850900950Wavelength (nm)0.0880.0900.0920.0940.0960.098Rp/Rs1xsolar No TiO VOA&A proofs: manuscript no. GTC_Wasp48
Lam, K. W. F., Faedi, F., Brown, D. J. A., et al. 2017, A&A, 599, A3
Lecavelier Des Etangs, A., Pont, F., Vidal-Madjar, A., & Sing, D. 2008, A&A,
481, L83
Murgas, F., Pallé, E., Zapatero Osorio, M. R., et al. 2014, A&A, 563, A41
Nortmann, L., Pallé, E., Murgas, F., et al. 2016, A&A, 594, A65
O'Rourke, J. G., Knutson, H. A., Zhao, M., et al. 2014, ApJ, 781, 109
Pallé, E., Chen, G., Alonso, R., et al. 2016, A&A, 589, A62
Palle, E., Chen, G., Prieto-Arranz, J., et al. 2017, A&A, 602, L15
Parviainen, H. 2015, MNRAS, 450, 3233
Parviainen, H. & Aigrain, S. 2015, MNRAS, 453, 3821
Parviainen, H., Pallé, E., Nortmann, L., et al. 2016, A&A, 585, A114
Pollacco, D. L., Skillen, I., Collier Cameron, A., et al. 2006, PASP, 118, 1407
Pont, F., Sing, D. K., Gibson, N. P., et al. 2013, MNRAS, 432, 2917
Redfield, S., Endl, M., Cochran, W. D., & Koesterke, L. 2008, ApJ, 673, L87
Rowe, J. F., Matthews, J. M., Seager, S., et al. 2006, ApJ, 646, 1241
Sada, P. V., Deming, D., Jennings, D. E., et al. 2012, PASP, 124, 212
Seager, S. & Sasselov, D. D. 1998, ApJ, 502, L157
Sharp, C. M. & Burrows, A. 2007, ApJS, 168, 140
Sing, D. K., Huitson, C. M., Lopez-Morales, M., et al. 2012, MNRAS, 426, 1663
Snellen, I. A. G., Albrecht, S., de Mooij, E. J. W., & Le Poole, R. S. 2008, A&A,
Turner, J. D., Pearson, K. A., Biddle, L. I., et al. 2016, MNRAS, 459, 789
Vidal-Madjar, A., Désert, J.-M., Lecavelier des Etangs, A., et al. 2004, ApJ, 604,
487, 357
L69
Fig. 8. Transmission spectrum of WASP-48b around the Na doublet.
The blue, red, and green lines show atmosphere models computed us-
ing Exo-Transmit for a WASP-48b analog with Teq = 2000 K, isother-
mal temperature profile, and Rayleigh scattering by H2 in the blue. The
dashed black line represent the Rp/Rs found for the white light curve
and ±3 atmospheric scale heights (H = 614.9 km) above and below this
level.
ing the wavelength region near the Na i doublet (λ 589.0 and
589.6 nm), we found no statistically significant detection of this
element.
Acknowledgements. Based on observations made with the Gran Telescopio Ca-
narias (GTC), installed in the Spanish Observatorio del Roque de los Muchachos
of the Instituto de Astrofísica de Canarias, in the island of La Palma. All the fig-
ures presented here were made using Matplotlib (Hunter 2007).
References
Ambikasaran, S., Foreman-Mackey, D., Greengard, L., Hogg, D. W., & O'Neil,
M. 2014, ArXiv e-prints [arXiv:1403.6015]
Bean, J. L., Désert, J.-M., Kabath, P., et al. 2011, ApJ, 743, 92
Bean, J. L., Désert, J.-M., Seifahrt, A., et al. 2013, ApJ, 771, 108
Burrows, A., Budaj, J., & Hubeny, I. 2008, ApJ, 678, 1436
Burrows, A., Hubeny, I., Budaj, J., Knutson, H. A., & Charbonneau, D. 2007,
ApJ, 668, L171
Burrows, A. S. 2014, Nature, 513, 345
Cepa, J., Aguiar, M., Escalera, V. G., et al. 2000, in Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series, Vol. 4008, Society
of Photo-Optical Instrumentation Engineers (SPIE) Conference Series, ed.
M. Iye & A. F. Moorwood, 623 -- 631
Charbonneau, D., Brown, T. M., Noyes, R. W., & Gilliland, R. L. 2002, ApJ,
Chen, G., Pallé, E., Nortmann, L., et al. 2017, A&A, 600, L11
Ciceri, S., Mancini, L., Southworth, J., et al. 2015, A&A, 577, A54
Désert, J.-M., Lecavelier des Etangs, A., Hébrard, G., et al. 2009, ApJ, 699, 478
Désert, J.-M., Vidal-Madjar, A., Lecavelier Des Etangs, A., et al. 2008, A&A,
Eastman, J., Siverd, R., & Gaudi, B. S. 2010, PASP, 122, 935
Enoch, B., Anderson, D. R., Barros, S. C. C., et al. 2011, AJ, 142, 86
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125,
Fortney, J. J., Lodders, K., Marley, M. S., & Freedman, R. S. 2008, ApJ, 678,
568, 377
492, 585
306
1419
Fortney, J. J., Shabram, M., Showman, A. P., et al. 2010, ApJ, 709, 1396
Giménez, A. 2006, A&A, 450, 1231
Hoeijmakers, H. J., de Kok, R. J., Snellen, I. A. G., et al. 2015, A&A, 575, A20
Hubeny, I., Burrows, A., & Sudarsky, D. 2003, ApJ, 594, 1011
Hunter, J. D. 2007, Computing In Science & Engineering, 9, 90
Husser, T.-O., Wende-von Berg, S., Dreizler, S., et al. 2013, A&A, 553, A6
Kempton, E. M.-R., Lupu, R. E., Owusu-Asare, A., Slough, P., & Cale, B. 2016,
ArXiv e-prints [arXiv:1611.03871]
Kreidberg, L., Bean, J. L., Désert, J.-M., et al. 2014, Nature, 505, 69
Article number, page 8 of 8
560570580590600610620630Wavelength (nm)0.0880.0900.0920.0940.0960.0980.100Rp/Rs1xsolar TiO VO1xsolar VO1xsolar No TiO VO |
1603.00414 | 1 | 1603 | 2016-03-01T19:19:09 | On the origin of the sub-Jovian desert in the orbital-period--planetary-mass plane | [
"astro-ph.EP"
] | Transit and radial velocity observations indicate a dearth of sub-Jupiter--mass planets on short-period orbits, outlined roughly by two oppositely sloped lines in the period--mass plane. We interpret this feature in terms of high-eccentricity migration of planets that arrive in the vicinity of the Roche limit, where their orbits are tidally circularized, long after the dispersal of their natal disk. We demonstrate that the two distinct segments of the boundary are a direct consequence of the different slopes of the empirical mass--radius relation for small and large planets, and show that this relation also fixes the mass coordinate of the intersection point. The period coordinate of this point, as well as the detailed shape of the lower boundary, can be reproduced with a plausible choice of a key parameter in the underlying migration model. The detailed shape of the upper boundary, on the other hand, is determined by the post-circularization tidal exchange of angular momentum with the star and can be reproduced with a stellar tidal quality factor $Q^\prime_*\sim10^6$. | astro-ph.EP | astro-ph | Draft version August 27, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
6
1
0
2
r
a
M
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
1
4
0
0
.
3
0
6
1
:
v
i
X
r
a
ON THE ORIGIN OF THE SUB-JOVIAN DESERT IN THE ORBITAL-PERIOD -- PLANETARY-MASS PLANE
Department of Astronomy & Astrophysics and The Enrico Fermi Institute, The University of Chicago, Chicago, IL 60637, USA
Titos Matsakos and Arieh Konigl
Draft version August 27, 2018
ABSTRACT
Transit and radial velocity observations indicate a dearth of sub-Jupiter -- mass planets on short-
period orbits, outlined roughly by two oppositely sloped lines in the period -- mass plane. We interpret
this feature in terms of high-eccentricity migration of planets that arrive in the vicinity of the Roche
limit, where their orbits are tidally circularized, long after the dispersal of their natal disk. We
demonstrate that the two distinct segments of the boundary are a direct consequence of the different
slopes of the empirical mass -- radius relation for small and large planets, and show that this relation
also fixes the mass coordinate of the intersection point. The period coordinate of this point, as well as
the detailed shape of the lower boundary, can be reproduced with a plausible choice of a key parameter
in the underlying migration model. The detailed shape of the upper boundary, on the other hand, is
determined by the post-circularization tidal exchange of angular momentum with the star and can be
reproduced with a stellar tidal quality factor Q′
Subject headings: planet-star interactions -- planets and satellites: dynamical evolution and stability
∗ ∼ 106.
-- planets and satellites: general
1. INTRODUCTION
As has been known for some time now, radial-velocity
surveys of exoplanets exhibit an abrupt drop in the num-
ber of hot Jupiters (HJs) for orbital periods Porb .
3 days (e.g., Cumming et al. 2008), corresponding to
a pileup of such planets near Porb = 3 days (e.g.,
Gaudi et al. 2005) and their paucity at shorter peri-
ods (e.g., Zucker & Mazeh 2002).
(We define HJs as
planets that have masses in the range Mp ∼ 0.3 -- 3 MJ,
where MJ is Jupiter's mass, and periods Porb . 10 days.)
Planet candidates identified by transit measurements
exhibit a similar sharp drop in the number count be-
low Porb ∼ 3 -- 4 days for objects with radii Rp & 4 R⊕
(e.g., Howard et al. 2012; Fressin et al. 2013). One in-
terpretation of these results invokes planet -- planet scat-
tering events that place one of the interacting planets
on a highly eccentric orbit: that planet thereby attains
a small pericenter distance, where its orbit is circular-
ized through tidal interaction with the host star (e.g.,
Ford & Rasio 2006). In this picture, the shortest possible
pericenter distance is given by the Roche limit aR, where
the planet starts to be tidally disrupted, and the planet's
resulting circular radius must therefore exceed ∼ 2 aR
(where the factor of 2 follows from conservation of orbital
angular momentum). A planet can also be placed on a
highly eccentric orbit by a more gradual process such as
Kozai (e.g., Wu & Murray 2003; Fabrycky & Tremaine
2007) or secular (e.g., Wu & Lithwick 2011; Petrovich
2015a) migration. An alternative interpretation at-
tributes the scarcity of high-Mp planets on short-period
orbits to tidal exchange of angular momentum with the
star, which causes close-in giant planets to spiral inward
and get ingested by their host (e.g., Jackson et al. 2009;
Teitler & Konigl 2014).
As data continued to accumulate, more details have
emerged about the structure of the planet distribution
at low values of Porb. In particular, Szab´o & Kiss (2011)
considered the (Porb, Mp) plane and identified a region
(defined by Porb < 2.5 days and Mp between 0.02 and
0.8 MJ) with a pronounced dearth of planets, which they
termed the "sub-Jupiter desert."1 They noted that this
region has both an upper boundary, consisting of HJs,
and a lower boundary, consisting of close-in super-Earths
(SEs), but is devoid of hot Neptunes. Similar identi-
fications were made by Beaug´e & Nesvorn´y (2013) and
by Mazeh et al. (2016) (with the latter authors approxi-
mating the boundary by two intersecting straight lines in
the (log Porb, log Mp) plane), and it is generally accepted
that this feature is not the result of some observational
bias.2 However, the origin of the desert is still being de-
bated (e.g., Szab´o & Kiss 2011; Ben´ıtez-Llambay et al.
2011; Beaug´e & Nesvorn´y 2013; Kurokawa & Nakamoto
2014; Batygin et al. 2016; Mazeh et al. 2016).
In one
proposed scenario (Kurokawa & Nakamoto 2014), the
desert is attributed to the evaporation of close-in giant
planets by the stellar radiation field, followed by Roche-
lobe overflow. However, this picture also implies the com-
plete loss of the gaseous envelopes of the less massive SEs,
which is inconsistent with the fact that the inferred radii
of the planets near the lower boundary of the desert are
generally well in excess of 1.6 R⊕ and are thus unlikely
to be purely rocky (see Rogers 2015). In an alternative
scenario (Valsecchi et al. 2014, 2015), HJs arrive at the
Roche limit and undergo rapid mass loss through Roche-
lobe overflow, with photoevaporation further contribut-
ing to the removal of their envelopes. It was proposed
that this mechanism could convert HJs into SEs, but it is
not obvious that the resulting low-mass planet distribu-
1 As discussed in that paper as well as in the other studies of
this topic that we cite, a complementary description of the sub-
Jovian desert can be given in the (Porb, Rp) plane. However, in
view of the degeneracy of the Mp(Rp) relation for large planets
(see Equation (1) below), we consider only the (Porb, Mp) plane
in this work.
2
In fact, several planets are known to lie inside the nominal
desert, although so far all such planets are located near the desert's
edges (e.g., Col´on et al. 2015).
2
Matsakos & Konigl
tion would be consistent with the observed shape of the
desert's lower boundary. Yet another possibility is that
the desert arises from the in-situ formation of HJs by gas
accretion onto SE cores (Boley et al. 2016; Batygin et al.
2016). However, this picture also faces various challenges
(e.g., Inamdar & Schlichting 2015).
In this Letter we propose that the boundary of the
desert reflects the locus of the innermost circularized or-
bits of planets that arrived in the vicinity of the host star
through planet-planet scattering, Kozai migration (aris-
ing from an interaction with either a stellar or a plane-
tary companion), or a secular process long after the pro-
toplanetary disk had dispersed. As was already pointed
out by Ford & Rasio (2006), the Roche limit aR (which is
taken to be the distance of closest approach for a planet
arriving via either scattering or Kozai migration) is a
function of both Rp and Mp (for a given host mass M∗),
and it can therefore be expressed (using the appropriate
mass -- radius relation) as a function of Mp. A similar ex-
pression can be written down for the expected distance
of closest approach in the secular chaos scenario. Using
an empirical Rp(Mp) relation, we demonstrate that this
picture can account for the basic shape of the desert's
boundary, with its oppositely sloped upper and lower
segments. We further show that the detailed structure
of the boundary can be qualitatively reproduced when
one also takes into account the orbital evolution induced
in close-in planets by their tidal interaction with the host
star. These results on the behavior of late-arriving close-
in planets complement recent inferences about the evo-
lution of an earlier generation of HJs that had migrated
though the natal disk (Matsakos & Konigl 2015).
2. MODELING APPROACH
pP 13/3
p , where Q′
orb MpR−5
Since most of the planets in the vicinity of the desert's
boundary are inferred to have low eccentricities (e < 0.1;
see Figure 1 below), we focus on scenarios in which plan-
ets are placed on high-eccentricity orbits that are cir-
cularized by internal tidal dissipation after the planets
arrive in the vicinity of the host star. For any given
system, we label by tarr the time at which the planet
under consideration arrives at the stellar vicinity with
orbital eccentricity e0. After arrival, its orbit is circular-
ized to a radius ac (where its orbital period is Porb) on a
timescale τc ∝ Q′
p is the planet's
tidal quality factor (e.g., Matsumura et al. 2010a). Plan-
ets along the lower boundary of the desert have low values
of Rp, which has the effect of increasing τc. However, the
smallest such planets would be mostly rocky and would
thus also have low values of Q′
p (e.g., Goldreich & Soter
1966), so their orbits might still be circularized over the
planets' typical ages (tage & 1 Gyr).
In this work we
relate Rp and Mp through the empirical relation ob-
tained by Weiss et al. (2013). They divided their sam-
ple into small and large planets (superscripts S and L,
respectively), separated at Mp ≈ 150 M⊕, and wrote
down power-law relationships that we further approxi-
mate here by adopting the median incident flux of the
data set (F = 8.6 × 108 erg s−1 cm−2) and rounding to
the first decimal figure in each exponent:
star
The
tidal
∗/Q′
interaction with the
can also
ingestion by
lead to orbital decay and eventual
This occurs on a nominal
timescale
the host.
p)(M∗/Mp)2(Rp/R∗)5] τc (where an aster-
τd ∼ [(Q′
isk again denotes the star), which is typically ≫ τc (e.g.,
Matsumura et al. 2010a). Given that τd ∝ 1/Mp, the
planets on the lower boundary of the desert are hardly
affected by this interaction; however, the initial period
distribution of the circularized orbits of large planets is
measurably modified.
2.1. Initial Locus of Innermost Planets with
Circularized Orbits
The post-circularization locus in the (Porb, Mp) plane
of the lowest-Porb planets can be obtained for both the
planet -- planet scattering and Kozai migration cases from
the expression for the Roche limit (RL),
aR = q (M∗/Mp)1/3 Rp ,
(2)
where the precise value of the coefficient q depends on
the planet's structural and orbital characteristics. Two
of the most frequently cited estimates are q = 2.16
(Paczy´nski 1971) and q = 2.7 (Guillochon et al. 2011),
but there are indications that its value in real systems
could be higher (up to ∼ 3.6 -- 3.8; Valsecchi & Rasio
2014; Petrovich 2015b). As it turns out, the interpre-
tation of the sub-Jovian desert in terms of the Roche
limit also favors a comparatively large value for q (≃ 3.5;
see Section 3). Under the assumption that e0 ≈ 1, the
circularization radius is ≃ 2 aR and thus can be written
as
ac,RL ≈ 0.03 (cid:16) q
3(cid:17)(cid:18) M∗
M⊙(cid:19)1/3(cid:18) Mp
MJ(cid:19)−1/3(cid:18) Rp
RJ(cid:19) AU .
(3)
In the secular chaos (SC) model, the circularization
radius was estimated by Wu & Lithwick (2011) as twice
the pericenter distance obtained by equating the preces-
sion rate of a planet's longitude of pericenter -- due to a
secular interaction with another planet located farther
out -- with the orbit-averaged precession rate due to the
tidal quadrupole induced on the planet by the star. This
yields
ac,SC ≈ 0.03(cid:18) α
× (cid:18) Mp
1/6(cid:19)−3/5(cid:18) M∗
MJ(cid:19)−1/5(cid:18) Rp
M⊙(cid:19)2/5(cid:18) Mpert
RJ(cid:19) AU ,
MJ (cid:19)−1/5
(4)
where Mpert is the mass of the perturbing planet and
α is the ratio of the semimajor axes of the two planets.
For the sake of simplicity, we henceforth fix α at its nor-
malization value and treat Mpert as the relevant model
parameter.3
By combining Equations (1), (3), and (4), and using
, we obtain the initial (post circularization)
Porb ∝ a−3/2
c
RS
p
RJ
∼ 1.6 (cid:18) Mp
MJ(cid:19)1/2
,
RL
p
RJ
MJ(cid:19)0
∼ 1.5 (cid:18) Mp
.
(1)
3 An alternative possibility would have been to treat the pa-
rameter combination α3Mpert as a variable.
Origin of Sub-Jovian Desert
3
TABLE 1
Model Parameters
Parameter
Sampled Value
Distribution
Number of planets
Porb [days]
Rp [R⊕]
for Rp < 12R⊕
Mp [M⊕] for Rp < 12R⊕
Rp [M⊕] for Rp > 12R⊕
Mp [M⊕] for Rp > 12R⊕
tarr [Gyr]
5
0.5 -- 50a
3 -- 12a
9 -- 144b
9 -- 20b
56 -- 5620b
0.01 -- 10
f (ln P ) ∝ P 0.47
f (ln R) ∝ R−0.66
(Rp/R⊕)2M⊕
Fit to data
Fit to data
Uniform in log time
References. -- a Youdin (2011); b Weiss et al. (2013).
locus of the innermost planets for these two scenarios,
orb,SC ∝ M −3/5
orb,SC ∝ M 3/20
P S
orb,RL ∝ M −1/2
orb,RL ∝ M 1/4
P S
RL :( P L
SC :( P L
p
p
p
p
,
.
(5)
It is seen that in both cases the inferred boundary has a
negatively sloped upper branch and a positively sloped
lower branch. This basic property of the observed sub-
Jovian desert is a generic feature of our proposed inter-
pretation, arising from the different slopes in the empir-
ical Rp(Mp) relation for large and small planets.
2.2. Effect of Tidally Induced Orbital Evolution
We calculate this effect following the procedure out-
lined in Matsakos & Konigl (2015). We perform Monte
Carlo simulations for a population of 30,000 plane-
tary systems, assuming solar-type hosts (M∗ = M⊙,
R∗ = R⊙) with initial rotation periods distributed uni-
formly in the range 5 -- 10 days, and drawing ages in
the range 1 -- 8 Gyr from the empirical distribution of
∗ = 106
Walkowicz & Basri (2013). We also adopt Q′
(see Section 3). The values of Porb and Rp are chosen
from the observationally inferred distributions presented
in Youdin (2011), with Mp deduced from the Rp(Mp)
compilation of Weiss et al. (2013) (using their inferred
power-law fit for small planets but accounting for the
observed scatter in both radius and mass for those with
Mp & 150 M⊕).4 We further assume a random distribu-
tion for the initial angle between the stellar spin and the
orbital plane, and that the distribution of planet arrival
times is uniform in log tarr (see Section 4). These choices
are summarized in Table 1.
For each system, we integrate the evolution equations
for the stellar and orbital angular momenta, taking into
account the effects of equilibrium tides and of magnetic
braking. We assume that all orbits in a multi-planet sys-
tem remain coplanar, but we neglect planet -- planet in-
teractions. We carry out the calculations for each of the
two distributions obtained in Section 2.1, removing from
consideration any planet drawn with an initial semima-
jor axis that is less than ac,RL (Equation (3)) or ac,SC
(Equation (4)). Given that we do not include systems
with tage ≤ 1 Gyr in our population counts, our pre-
4 The adopted distributions of Porb and Rp are in the form of
a single power law and thus do not capture the detailed behavior
of planets in this region of the period -- mass plane. However, since
our focus is on the shape of the desert's boundary and not on
reproducing the observed density of planets, we do not consider
more elaborate distributions.
dicted distributions do not reveal the possible presence
of close-in giant planets that arrive at the stellar vicin-
ity by migration through the natal disk. The contribu-
tion of such early-arriving planets to the observed num-
ber count is, however, small due to their relatively rapid
tidal ingestion by the star. Nevertheless, as was noted
by Matsakos & Konigl (2015), these HJs could have a
strong influence on the observed distribution of the an-
gle between the stellar spin and the orbital plane. Based
on that work, we take account of this effect by including,
at time t = 0, a "stranded HJ" (SHJ) characterized by
MSHJ = 0.6 MJ, RSHJ = RJ, and PSHJ = 2 days in (ran-
domly selected) 50% of the modeled systems. A giant
planet of this type would undergo tidal ingestion by a
G-type star on a timescale of ∼ 0.7 Gyr.
3. RESULTS
Figure 1 presents the predicted distributions in the
(Porb, log Mp) plane for the RL and SC models. For
the RL model we also plot the relation obtained from
combining Equations (1) and (3) for three values of the
parameter q,5 whereas for the SC model we show the cor-
responding relation obtained from Equations (1) and (4)
for three values of Mpert.6 It is seen that each of these
scenarios can reproduce the observed shape of the sub-
Jovian desert's boundary quite well. The initial circular-
ization locus accounts for the basic "bird's beak" configu-
ration of the boundary, but only the shape of its low-mass
segment (along which orbital evolution effects are negli-
gible) and the location of the vertex (where the upper
and lower segments intersect) are preserved in the final
distribution. The mass coordinate of the vertex is fixed
by the transition point (Mp ≈ 150 M⊕) of the empirical
mass -- radius relation and is independent of the details of
the underlying planet migration model; the fact that it
matches the observations so accurately is a strong indica-
tion that this aspect of the small/large planets dichotomy
plays a key role in determining the desert's shape. On
the other hand, the plots in the left panels of Figure 1
demonstrate that the period coordinate of the vertex is
sensitive to the value of the relevant model parameter. In
each case, however, the inferred best-fit value has a plau-
sible magnitude. In the RL picture, the comparatively
large indicated value of q (≃ 3.5) is consistent with other
recent determinations (e.g., Valsecchi & Rasio 2014, who
derived the initial locations of HJs detected inside the
Roche limit; and Petrovich 2015b, who investigated the
formation of HJs by Kozai migration).
In the case of
secular chaos, the inference that Mpert ≈ Mp is also em-
inently reasonable. For Jupiter-mass planets, the exci-
tation of secular interactions would be much weaker if
the companion planet's mass were measurably less than
that of the innermost planet (e.g., Wu & Lithwick 2011),
whereas for SE planets, comparable-mass objects are the
most frequent companions and are thus the most likely
to act as perturbers. Note that the value of the model
parameter (q or Mpert) also affects the predicted shape of
5 The values 2.16 and 2.70 represent the estimates listed in
Section 2.1, whereas q = 3.46 corresponds to our best fit to the
desert's shape.
6 In deriving these relations, we use the power-law scalings listed
in Equation (1) for small and large planets, and divide the two
populations at the value of Mp (≃ 150 M⊕) given in Weiss et al.
(2013).
4
]
J
M
[
s
s
a
m
t
e
n
a
p
l
]
J
M
[
s
s
a
m
t
e
n
a
p
l
Matsakos & Konigl
ROCHE LIMIT MODEL
model data, initial
Eqs. (1) & (3): q =3.46
Eqs. (1) & (3): q =2.70
Eqs. (1) & (3): q =2.16
]
J
M
[
s
s
a
m
t
e
n
a
p
l
ROCHE LIMIT MODEL
model data, final
obs. data: e < 0.1
obs. data: e > 0.1
obs. data: e n/a
sub-Jovian desert
101
100
10-1
101
100
10-1
0.5
1.0
1.5
2.5
2.0
3.5
orbital period [days]
3.0
4.0
4.5
5.0
0.5
1.0
1.5
2.5
2.0
3.5
orbital period [days]
3.0
4.0
4.5
5.0
model data, initial
Eqs. (1) & (4): Mpert =Mp
Eqs. (1) & (4): Mpert =2Mp
Eqs. (1) & (4): Mpert =MJ/2
SECULAR CHAOS
MODEL
]
J
M
[
s
s
a
m
t
e
n
a
p
l
SECULAR CHAOS MODEL
model data, final
obs. data: e < 0.1
obs. data: e > 0.1
obs. data: e n/a
sub-Jovian desert
101
100
10-1
101
100
10-1
0.5
1.0
1.5
2.5
2.0
3.5
orbital period [days]
3.0
4.0
4.5
5.0
0.5
1.0
1.5
2.5
2.0
3.5
orbital period [days]
3.0
4.0
4.5
5.0
Fig. 1. -- Predicted vs. observed planet distributions in the (Porb, log Mp) plane. The top panels correspond to the Roche limit model for
the innermost circularized orbits (applicable to planets arriving at the stellar vicinity through either scattering or Kozai migration), whereas
the bottom panels correspond to the secular migration model. For each of these two cases, the left and right panels show, respectively,
the initial distribution of planets on circular orbits and the final distribution obtained by calculating the effect of tidally induced orbital
evolution. The initial planet distribution in the left panels was obtained by sampling empirical distributions of the orbital period, planetary
radius, and system age as well as the adopted arrival-time distribution and the empirical Rp(Mp) relation (see Table 1). Analytic predictions
for the inner edge of this distribution are also shown in each case in the left panel for several values of the relevant model parameter. The
data shown in the right panels are for all confirmed planets listed in exolanets.org (Han et al. 2014) as of 2015 December 16, and are
color coded according to their eccentricity values (when available). To improve the presentation, the number of displayed low-mass model
planets was randomly reduced by 95% in each of the panels.
Origin of Sub-Jovian Desert
5
the lower branch of the desert's boundary, which further
constrains the choice of the best match.
The shape of the upper boundary is largely indepen-
dent of the details of the initial distribution of high-
mass planets, and is determined mainly by the ensu-
ing orbital evolution. It is well reproduced -- in the con-
text of the equilibrium tidal interaction model that we
∗ ∼ 106,
employ -- with a stellar tidal quality factor Q′
which is consistent with previous inferences from mod-
eling the Porb distribution of HJs (e.g., Teitler & Konigl
2014; Essick & Weinberg 2016).
It is noteworthy that
the two alternative models originally proposed to explain
the spatial distribution of HJs -- circularization of highly
eccentric orbits and tidal exchange of angular momen-
tum with the star -- are both found to be relevant in the
context of the current interpretation when the detailed
distribution in the (Porb, Mp) plane is taken into account.
4. DISCUSSION
Our proposed interpretation of the sub-Jovian desert's
shape -- that it is a natural consequence of the orbital
circularization of planets that arrive at the stellar vicin-
ity on high-eccentricity orbits and of their subsequent
tidal angular-momentum exchange with the star -- is con-
sistent with the finding that the eccentricity distri-
bution of giant planets broadens with increasing pe-
riod (e.g., Winn & Fabrycky 2015) and with the evi-
dence for ongoing orbital evolution of the innermost
HJs (e.g., Valsecchi & Rasio 2014; Teitler & Konigl 2014;
Essick & Weinberg 2016). The latter result, in turn, sup-
ports the view that the observed HJs are mostly late-
arriving planets rather than the product of migration in
the protoplanetary disk. The high-eccentricity migra-
tion scenario, in which the ingoing planet is placed on a
high-e0 orbit through gravitational interaction with one
or more massive bodies (other planets or a binary star),
is consistent with the indicated high occurrence rate of
planetary and stellar companions in systems that harbor
HJs (e.g. Knutson et al. 2014; Ngo et al. 2015). We con-
sidered three possible pathways to such an outcome --
planet-planet scattering, Kozai migration, and secular
drift -- and inferred that in principle they could all play
a role. However, a more detailed examination is required
to determine the actual contribution of each of these pro-
cesses. One pertinent question is whether the influence
of the process extends to late times; for example, sec-
ular chaos is inherently a long-lasting interaction, for
which the expected distribution of arrival times is ap-
proximately uniform in log tarr (Y. Lithwick, personal
communication), whereas scattering likely only plays a
role at early times (. 108 yr; e.g., Chatterjee et al. 2008;
Juri´c & Tremaine 2008). Another relevant question con-
cerns the rate of eccentricity growth in the high-e migra-
tion model; for example, Wu & Lithwick (2011) argued
that the observed ∼ 3-day pileup of HJs is more likely
to have been caused by a slow process such as Kozai or
secular migration than by sudden scattering events (see
also Nagasawa et al. 2008). Further studies are needed
to fully address this issue.
One attractive feature of the proposed interpretation
is that it can simultaneously reproduce both the up-
per and the lower segments of the desert's boundary.
An alternative possibility for the low-mass branch is
that it arises from the in-situ formation of close-in SEs
(Mazeh et al. 2016), which has been recently explored
in the literature (e.g., Lee & Chiang 2016). However,
arguments have also been given in favor of formation at
larger distances (up to a few AU) and subsequent inward
migration (e.g., Schlichting 2014; Inamdar & Schlichting
2015). Our model for the sub-Jovian desert is consistent
with the latter picture.
The scenario considered in this work, which involves
late-arriving planets, complements the SHJs model pre-
sented in Matsakos & Konigl (2015), which is concerned
with early-arriving HJs. That model explains the good
alignment exhibited by a significant fraction of HJs
around cool (G type) stars -- as well as the good align-
ment inferred for more distant planets around such
stars -- vs.
the broad range of obliquities exhibited by
HJs around hot (F type) stars. As we noted in Sec-
tion 2, the earlier population of HJs would not affect the
observed properties of the sub-Jovian desert because of
the relatively short SHJ ingestion time. However, consis-
tency with the SHJs scenario requires the orbital plane
of any late-arriving planet to roughly coincide with that
of the natal disk. This additional constraint on the high-
eccentricity migration model need not, however, be too
difficult to fulfill. For example, Matsumura et al. (2010b)
found that only ∼ 15% of planets in a 3-planet system
that emerges out of a gas disk have orbital inclinations
> 10◦, and Lithwick & Wu (2014) discovered that, even
if larger initial inclinations are allowed, 60% of the HJs
formed through secular migration involving a 3-planet
system have projected obliquities < 10◦. Furthermore,
Petrovich (2015a) demonstrated the feasibility of pro-
ducing HJs through coplanar high-e migration (a secular
process involving 2 planets).
We are grateful to Tsevi Mazeh for alerting us to the
"bird's beak" shape of the sub-Jovian desert's boundary.
We thank him as well as Dan Fabrycky, Yoram Lith-
wick, Leslie Rogers, and the referee for helpful input.
This work was supported in part by NASA ATP grant
NNX13AH56G.
REFERENCES
Batygin, K., Bodenheimer, P. H., & Laughlin, G. P. 2016, ApJ,
Col´on, K. D., Morehead, R. C., & Ford, E. B. 2015, MNRAS,
submitted, arXiv:1511.09157
Beaug´e, C., & Nesvorn´y, D. 2013, ApJ, 763, 12
Ben´ıtez-Llambay, P., Masset, F., & Beaug´e, C. 2011, A&A, 528,
A2
Boley, A. C., Granados Contreras, A. P., & Gladman, B. 2016,
ApJ, 817, L17
Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008,
ApJ, 686, 580
452, 3001
Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008, PASP,
120, 531
Essick, R., & Weinberg, N. N. 2016, ApJ, 816, 18
Fabrycky, D., & Tremaine, S. 2007, ApJ, 669, 1298
Ford, E. B., & Rasio, F. A. 2006, ApJ, 638, L45
Fressin, F., Torres, G., Charbonneau, D., et al. 2013, ApJ, 766, 81
6
Matsakos & Konigl
Gaudi, B. S., Seager, S., & Mallen-Ornelas, G. 2005, ApJ, 623,
472
Goldreich, P., & Soter, S. 1966, Icarus, 5, 375
Guillochon, J., Ramirez-Ruiz, E., & Lin, D. 2011, ApJ, 732, 74
Han, E., Wang, S. X., Wright, J. T., et al. 2014, PASP, 126, 827
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS,
201, 15
Inamdar, N. K., & Schlichting, H. E. 2015, MNRAS, 448, 1751
Jackson, B., Barnes, R., & Greenberg, R. 2009, ApJ, 698, 1357
Juri´c, M., & Tremaine, S. 2008, ApJ, 686, 603
Knutson, H. A., Fulton, B. J., Montet, B. T., et al. 2014, ApJ,
785, 126
Kurokawa, H., & Nakamoto, T. 2014, ApJ, 783, 54
Lee, E. J., & Chiang, E. 2016, ApJ, 817, 90
Lithwick, Y., & Wu, Y. 2014, Proceedings of the National
Academy of Science, 111, 12610
Matsakos, T., & Konigl, A. 2015, ApJ, 809, L20
Matsumura, S., Peale, S. J., & Rasio, F. A. 2010a, ApJ, 725, 1995
Matsumura, S., Thommes, E. W., Chatterjee, S., & Rasio, F. A.
2010b, ApJ, 714, 194
Mazeh, T., Holczer, T., & Faigler, S. 2016, A&A, submitted
Nagasawa, M., Ida, S., & Bessho, T. 2008, ApJ, 678, 498
Ngo, H., Knutson, H. A., Hinkley, S., et al. 2015, ApJ, 800, 138
Paczy´nski, B. 1971, ARA&A, 9, 183
Petrovich, C. 2015a, ApJ, 805, 75
-- . 2015b, ApJ, 799, 27
Rogers, L. A. 2015, ApJ, 801, 41
Schlichting, H. E. 2014, ApJ, 795, L15
Szab´o, G. M., & Kiss, L. L. 2011, ApJ, 727, L44
Teitler, S., & Konigl, A. 2014, ApJ, 786, 139
Valsecchi, F., Rappaport, S., Rasio, F. A., Marchant, P., &
Rogers, L. A. 2015, ApJ, 813, 101
Valsecchi, F., & Rasio, F. A. 2014, ApJ, 787, L9
Valsecchi, F., Rasio, F. A., & Steffen, J. H. 2014, ApJ, 793, L3
Walkowicz, L. M., & Basri, G. S. 2013, MNRAS, 436, 1883
Weiss, L. M., Marcy, G. W., Rowe, J. F., et al. 2013, ApJ, 768, 14
Winn, J. N., & Fabrycky, D. C. 2015, ARA&A, 53, 409
Wu, Y., & Lithwick, Y. 2011, ApJ, 735, 109
Wu, Y., & Murray, N. 2003, ApJ, 589, 605
Youdin, A. N. 2011, ApJ, 742, 38
Zucker, S., & Mazeh, T. 2002, ApJ, 568, L113
|
1610.07632 | 1 | 1610 | 2016-10-24T20:12:23 | Forward and Inverse Modeling of the Emission and Transmission Spectrum of GJ 436b: Investigating Metal Enrichment, Tidal Heating, and Clouds | [
"astro-ph.EP"
] | The Neptune-mass GJ 436b is one of the most-studied transiting exoplanets with repeated measurements of both its thermal emission and transmission spectra. We build on previous studies to answer outstanding questions about this planet, including its potentially high metallicity and tidal heating of its interior. We present new observations of GJ 436b's thermal emission at 3.6 and 4.5 micron, which reduce uncertainties in estimates of GJ 436b's flux at those wavelengths and demonstrate consistency between Spitzer observations spanning more than 7 years. We analyze the Spitzer thermal emission photometry and Hubble WFC3 transmission spectrum in tandem. We use a powerful dual-pronged modeling approach, comparing these data to both self-consistent and retrieval models. We vary the metallicity, intrinsic luminosity from tidal heating, disequilibrium chemistry, and heat redistribution. We also study the effect of clouds and photochemical hazes on the spectra, but do not find strong evidence for either. The self-consistent and retrieval modeling combine to suggest that GJ 436b has a high atmospheric metallicity, with best fits at or above several hundred times solar metallicity, tidal heating warming its interior with best-fit intrinsic effective effective temperatures around 300--350 K, and disequilibrium chemistry. High metal-enrichments (>600x solar) can only occur from the accretion of rocky, rather than icy, material. Assuming Tint~300--350 K, we find that Q'~2x10^5--10^6, larger than Neptune's Q', and implying a long tidal circularization timescale for the planet's orbit. We suggest that Neptune-mass planets may be a more diverse class than previously imagined, with metal-enhancements potentially spanning several orders of magnitude, to perhaps over 1000x solar metallicity. High fidelity observations with instruments like JWST will be critical for characterizing this diversity. | astro-ph.EP | astro-ph |
DRAFT FOR APJ
Preprint typeset using LATEX style emulateapj v. 5/2/11
FORWARD AND INVERSE MODELING OF THE EMISSION AND TRANSMISSION SPECTRUM OF GJ 436B:
INVESTIGATING METAL ENRICHMENT, TIDAL HEATING, AND CLOUDS
CAROLINE V. MORLEY1, 2, HEATHER KNUTSON3, MICHAEL LINE4, JONATHAN J. FORTNEY5, DANIEL THORNGREN6, MARK S.
MARLEY7, DILLON TEAL5, ROXANA LUPU7
Draft for ApJ
ABSTRACT
The Neptune-mass GJ 436b is one of the most-studied transiting exoplanets with repeated measurements
of both its thermal emission and transmission spectra. We build on previous studies to answer outstanding
questions about this planet, including its potentially high metallicity and tidal heating of its interior. We present
new observations of GJ 436b's thermal emission at 3.6 and 4.5 µm, which reduce uncertainties in estimates of
GJ 436b's flux at those wavelengths and demonstrate consistency between Spitzer observations spanning more
than 7 years. We analyze the Spitzer thermal emission photometry and Hubble WFC3 transmission spectrum
in tandem. We use a powerful dual-pronged modeling approach, comparing these data to both self-consistent
and retrieval models. We vary the metallicity, intrinsic luminosity from tidal heating, disequilibrium chemistry,
and heat redistribution. We also study the effect of clouds and photochemical hazes on the spectra, but do
not find strong evidence for either. The self-consistent and retrieval modeling combine to suggest that GJ
436b has a high atmospheric metallicity, with best fits at or above several hundred times solar metallicity,
tidal heating warming its interior with best-fit intrinsic effective effective temperatures around 300–350 K, and
disequilibrium chemistry. High metal-enrichments (> 600× solar) can only occur from the accretion of rocky,
rather than icy, material. Assuming Tint∼300–350 K, we find that Q(cid:48) ∼ 2× 105–106, larger than Neptune's Q(cid:48),
and implying a long tidal circularization timescale for the planet's orbit. We suggest that Neptune-mass planets
may be a more diverse class than previously imagined, with metal-enhancements potentially spanning several
orders of magnitude, to perhaps over 1000× solar metallicity. High fidelity observations with instruments like
JWST will be critical for characterizing this diversity.
Subject headings: keywords
1. INTRODUCTION
Determining the compositions of exoplanets ranging from
Earth-mass to Jupiter-mass in different environments is a key
goal of exoplanetary research. Planetary compositions are
shaped by the details of planet formation and altered by at-
mospheric physics and chemistry. Over a decade after its dis-
covery by Butler et al. (2004), GJ 436b remains the planet in
its Neptune-mass class for which we have obtained the most
detailed observations of its atmosphere.
GJ 436b was discovered to transit by Gillon et al. (2007b)
and, as the smallest transiting planet in 2007 and a favorable
target for observations, immediately became a target for atmo-
spheric characterization studies with the Spitzer Space Tele-
scope and Hubble Space Telescope. It remains one of the most
favorable and interesting targets for followup spectroscopic
studies: to date a total of 18 secondary eclipses and 8 transits
have been observed with Spitzer, along with 7 transits with
HST (Deming et al. 2007; Demory et al. 2007; Gillon et al.
2007a; Stevenson et al. 2010; Beaulieu et al. 2011; Knutson
et al. 2011, 2014a).
The atmosphere of GJ 436b has been a perennial challenge
1 Department of Astronomy, Harvard University, Cambridge, MA
02138; [email protected]
6 Department of Physics, University of California, Santa Cruz
7 NASA Ames Research Center
2 NASA Sagan Fellow
3 Division of Geological and Planetary Sciences, California Institute of
4 School of Earth and Space Exploration, Arizona State University
5 Department of Astronomy and Astrophysics, University of California,
Technology
Santa Cruz
to understand. Previous observations and modeling efforts,
which we describe below, have suggested high metallicity
compositions with strong vertical mixing. Many of these con-
clusions rest on the robustness of the Spitzer 3.6 and 4.5 µm
eclipses. Here, we move forward to study this planet us-
ing both its thermal emission photometry and its transmission
spectrum, adding three new eclipse observations at these two
wavelengths and analyzing the dataset with a powerful dual-
pronged approach of self-consistent and retrieval modeling.
1.1. Observations and Interpretation of Thermal Emission
Secondary eclipse measurements allow us to infer the
planet's brightness, and therefore temperature, as a function
of wavelength when the planet passes behind the host star. A
planet will appear fainter, and therefore create a shallower oc-
cultation, at wavelengths of strong absorption features, and
appear brighter at wavelengths of emission features.
The first secondary eclipse measurements of GJ 436b were
observed at 8 µm, while Spitzer was still operating cryogeni-
cally (Deming et al. 2007; Demory et al. 2007). These obser-
vations revealed that GJ 436b has a high eccentricity, ∼0.15,
which, given predicted tidal circularization timescales, sug-
gests the presence of a companion and of potential tidal heat-
ing (Ribas et al. 2008; Batygin et al. 2009).
With an equilibrium temperature around 700–800 K, GJ
436b is cool enough that models assuming thermochemical
equilibrium predict high CH4 abundance and low CO and CO2
abundance, which would result in a deeper occultation at 4.5
µm than 3.6 µm. However, when Stevenson et al. (2010) pub-
lished the first multi-wavelength thermal emission spectrum
of GJ 436b, measuring photometric points at 3.6, 4.5, 5.8, 8.0,
2
16, and 24 µm, they found that its occultation was deeper at
3.6 µm and shallower at 4.5 µm and suggested methane deple-
tion due to photo-dissociation as an explanation. Additional
studies have reanalyzed these observations and observed ad-
ditional secondary eclipses (Knutson et al. 2011; Lanotte et al.
2014).
In particular, the analysis by Lanotte et al. (2014)
revealed a significantly shallower 3.6 µm eclipse and some-
what shallower 8.0 µm eclipse; no detailed atmospheric stud-
ies have been carried out since these revisions.
From the time of the initial observations of GJ 436b's ther-
mal emission, it has been a major challenge to find self-
consistent models that adequately explain the data. Mad-
husudhan & Seager (2011) found using retrieval algorithms
that the atmosphere is best fit by an atmosphere rich in CO
and CO2 and depleted in CH4. Line et al. (2011) used dis-
equilibrium chemical models including the effect of photo-
chemistry, but found that they were not able to reproduce the
low observed methane abundance. Moses et al. (2013) found
that high metallicities (230–1000× solar) favor the high CO
and CO2 abundances inferred from the observations. Agún-
dez et al. (2014), noting the high eccentricity of GJ 436b,
studied the effect of tidal heating deep in the atmosphere on
the chemistry and find that significant tidal heating and high
metallicities fit the observed photometry best.
1.2. Observations and Interpretation of Transmission
Spectrum
Wavelength-dependent observations of the transit depth of
GJ 436b allow us to probe the composition of GJ 436b's day–
night terminator. At wavelengths with strong absorption fea-
tures, the planet will occult a larger area of the star, result-
ing in a deeper transit depth. Pont et al. (2009) observed the
transmission spectrum of GJ 436b from 1.1 to 1.9 µm with
NICMOS on HST but due to systematic effects were unable
to achieve high enough precision to detect the predicted water
vapor feature. Beaulieu et al. (2011) presented transit mea-
surements in Spitzer's 3.6, 4.5, and 8.0 µm filters that showed
higher transit depths at 3.6 and 8.0 µm than at 4.5 µm, indi-
cating strong methane absorption. Knutson et al. (2011) ana-
lyzed the same data and suggested that variable stellar activity
caused the observed transit depth at 3.6 µm to vary between
epochs. However, these data were reanalyzed again by Lan-
otte et al. (2014) and Morello et al. (2015) with new tech-
niques, which both found that the transit depths were constant
in the different bandpasses and remained constant between
epochs of observations.
More recently, Knutson et al. (2014a) used WFC3 on HST
to measure the transmission spectrum from 1.1–1.7 µm. Like
Pont et al. (2009), they do not detect a water vapor feature, but
with their higher S/N spectrum are able to rule out a cloud-
free H/He-dominated atmosphere to high confidence (48σ).
The spectrum is consistent with a high cloud at pressures of
∼1 mbar, or a H/He poor (3% H/He by mass, 1900× solar)
atmospheric composition.
1.3. A Third Body in the GJ 436 System?
Because the tidal circularization timescales are predicted
to be shorter than the age of the star, the non-zero eccen-
tricity of GJ 436b has suggested that it may have at least
one companion in the system; however, a number of searches
for additional planets did not find additional bodies (Demory
et al. 2007; Deming et al. 2007; Maness et al. 2007; Alonso
et al. 2008; Ribas et al. 2008; Cáceres et al. 2009; Ballard
et al. 2010b,a; Beust et al. 2012). Stevenson et al. (2012) an-
nounced the detection of two candidate sub-Earth-size planets
in the system, but later work by Lanotte et al. (2014) did not
find any evidence of these candidate companions.
1.4. The Need for an Additional Atmospheric Study
Is it
Here, we build on this extensive history of both observa-
tions and modeling for this enigmatic warm Neptune to an-
swer the still-outstanding questions about this planet. Do
the revisions in the eclipse points from Lanotte et al. (2014)
change the inferred composition?
truly ultra-high
(>300× solar) metallicity? What atmospheric physics must
be present for a Neptune-mass planet to have the observed
spectra and inferred atmospheric composition?
To these ends, we present an additional three secondary
eclipse observations (1 at 3.6 µm, 2 at 4.5 µm), demonstrat-
ing the robustness of these observations with modern Spitzer
observational and analysis techniques. We study both the ther-
mal emission and transmission spectra of GJ 436b in tandem,
including the published dataset of Spitzer photometry span-
ning from 3.6 to 16 µm and the transmission spectrum from
HST/WFC3. Unlike most previous studies, we investigate
whether including clouds or hazes in GJ 436b's atmosphere
can match both sets of observations for Neptune-like compo-
sitions (50–300 × solar), without invoking ultra-high metal-
licity (>1000× solar) compositions. We combine our self-
consistent treatment with results from chemically-consistent
retrievals that do not include clouds, and show that H/He-poor
atmospheric compositions with tidal heating provide the most
precise fit to GJ 436b's thermal emission spectrum, while also
fitting the transmission spectrum.
1.4.1. Format of this work
In Section 2 we describe the observations and data analysis.
In Section 3, we describe the modeling tools used to simulate
the observations, including both self-consistent and retrieval
models. In Section 4.2 we compare the data to self-consistent
models; in Section 4.3 we use retrieval algorithms to retrieve
chemical abundances and pressure–temperature profile and
compare these results with the results from self-consistent
modeling.
2. OBSERVATIONS AND DATA ANALYSIS
2.1. Photometry and Instrumental Model
These observations were obtained in the 3.6 and 4.5 µm
bandpasses using the Infra-Red Array Camera (IRAC) on
the Spitzer Space Telescope. In this paper we present three
new secondary eclipse observations of this planet, including
a 3.6 µm observation obtained on UT 2014 Jul 29 and two
4.5 µm observations obtained on UT 2014 Aug 11 and UT
2015 Feb 25, respectively, as part of Spitzer program 50056
(PI: Knutson). We also re-examine three archival eclipse ob-
servations including a 3.6 µm eclipse from UT 2008 Jan 30, as
well as 4.5 µm eclipses from UT 2008 Feb 2 and UT 2011 Jan
24 (Stevenson et al. 2010, 2012; Lanotte et al. 2014). Eclipses
from 2008 were observed during Spitzer's cryogenic mission,
while the remaining eclipses were observed during the ex-
tended warm mission. All eclipses were observed in subarray
mode, with integration times and observation durations given
in Table 1. Our new 2014-2015 observations included a now-
standard 30-minute peak-up pointing observation prior to the
start of our science observations. This adjustment corrects the
initial telescope pointing in order to place the star near the
GJ 436b
TABLE 1
Spitzer OBSERVATION DETAILS
3
c
c
c
c Bkd (%)e
a
tint (s)b
Length (h)
5.9
4.5
5.9
6.1
4.5
4.5
nimg
163,200
122,112
49,920
51,712
122,112
122,112
λ (µm) UT Start Date
2008-01-30
3.6
3.6
2014-07-29
2008-02-02
4.5
2011-01-24
4.5
2014-08-11
4.5
4.5
2015-02-25
a Total number of images.
b Integration time.
c ttrim is the amount of time in hours trimmed from the start of each time series, nbin is the bin size used in the photometric
fits, rpos is the radius of the aperture used to determine the position of the star on the array, and rphot is the radius of the
photometric aperture in pixels.
e Sky background contribution to the total flux for the selected aperture.
rphot
2.8
2.5
2.9
4.5
2.7
2.8
nbin
192
128
32
32
128
128
ttrim
1.0
1.0
3.0
2.0
0.5
0.5
rpos
3.0
2.0
2.0
2.0
2.0
2.0
0.05
0.25
0.09
0.38
0.11
0.12
0.1
0.1
0.4
0.4
0.1
0.1
center of the pixel where the effect of intrapixel sensitivity
variations is minimized.
We utilize BCD image files for our photometric analysis
and extract BJDUTC mid-exposure times using the information
in the image headers. The sky background for each 32×32
pixel image is calculated by excluding a circular region with
a radius of twelve pixels centered on the star, taking all of
the remaining pixels and trimming outliers greater than three
standard deviations away from the median, and then fitting a
Gaussian function to a histogram of the remaining pixels. We
calculate the flux-weighted centroid position of the star on the
array and derive the corresponding total flux in a circular aper-
ture for each individual image as described in previous stud-
ies (e.g. Lewis et al. 2013; Deming et al. 2015; Kammer et al.
2015). We consider both fixed and time varying photometric
aperture sizes in our fits but find that in all cases we obtain
a lower RMS and reduced levels of time-correlated ("red")
noise in our best-fit residuals using fixed apertures, in good
agreement with the conclusions of Lanotte et al. (2014). We
consider apertures with radii ranging between 2.0−5.0 pixels,
where we step in increments of 0.1 pixels between 2.0 − 3.0
pixels and in 0.5 pixel increments for larger radii.
The sensitivity of individual 3.6 and 4.5 µm IRAC pixels
varies from the center to the edge; when combined with short-
term telescope pointing oscillations, this produces variations
in the raw stellar fluxes plotted in Fig. 1. We correct for
this effect using the pixel-level decorrelation (PLD) method
(Deming et al. 2015), which produces results that are compa-
rable to or superior to those from a simple polynomial decor-
relation or pixel mapping method for light curves with du-
rations of less than ten hours (for a discussion of the PLD
method applied to longer phase curve observations, see Wong
et al. 2015). We utilize the raw flux values in a 3× 3 grid of
pixels centered on the position of the star, and then normal-
ize these individual pixel values by dividing by the total flux
in each 3× 3 postage stamp. We then incorporate these light
curves into an instrumental model given by:
(cid:80)
wiFi(t)(cid:80)
Fi(t)
i
i
Fmodel(t) =
(1)
where Fmodel is the predicted stellar flux in an individual im-
age, Fi is the measured flux in the ith individual pixel, and
wi is the weight associated with that pixel. We leave these
weights as free parameters in our fit, and solve for the values
that best match our observed light curves simultaneously with
FIG. 1.- Raw Spitzer 3.6 and 4.5 µm photometry as a function of time
from the center of eclipse phase reported in Knutson et al. (2011). We bin
the photometry in 30 s (grey filled circles) and 5 minute (black filled circles)
intervals, and overplot the best fit instrumental models binned in 5 minute
intervals for comparison (solid lines).
our eclipse fits.
Following the example of Deming et al. (2015), we fit this
model to binned light curves with optimal bin sizes given in
Table 1, where we considered a range of bin sizes between 1–
512 points sampled in steps of 2n. After identifying the best-
fit model we apply this solution to the unbinned light curves
in order to generate a corresponding vector of unbinned resid-
uals, which we use to evaluate the noise properties of the
data. As discussed in Deming et al. (2015) and Kammer et al.
(2015), we create a metric to measure the noise properties of a
given version of the photometry by calculating the root mean
square (RMS) variance of the residuals as a function of bin
size (Fig. 2). We then take the difference between a Gaussian
0.9900.9951.0001.0051.0100.9900.9951.0001.0051.010-4-3-2-101230.9900.9951.0001.0051.010-4-3-2-10123Time From Predicted Eclipse Center [h]Relative Flux3.6 um, UT 2008 Jan 303.6 um, UT 2014 Jul 294.5 um, UT 2008 Feb 24.5 um, UT 2011 Jan 244.5 um, UT 2014 Aug 114.5 um, UT 2015 Feb 254
FIG. 2.- Standard deviation of the best-fit residuals as a function of the
number of data points per bin
n scaling for Gaussian noise as
(black lines). We over plot the expected 1/
red dashed lines, where we have normalized these lines to match the standard
deviation of the unbinned residuals.
√
√
n scaling and the observed RMS as a
noise model with 1/
function of bin size, square the difference, and sum over all
bins. We then pick the version of the photometry that has the
lowest amount of red noise as measured by our least squares
metric after discarding solutions where the RMS of the best-fit
residuals is more than 1.1 times higher than the lowest RMS
version of the photometry. We trim a small section of data
from the start of each light curve in order to remove the ex-
ponential ramp, which is another well-known feature of the
IRAC 3.6 and 4.5 µm arrays (e.g., Lewis et al. 2013; Zellem
et al. 2014). After selecting the optimal aperture and bin sizes,
we examine the normalized light curves after detector effects
have been removed and vary the trim duration until we iden-
tify the value which minimizes the red noise in the best-fit
residuals, in agreement with our criteria for selecting the op-
timal aperture and bin size. Because we do not include an ex-
ponential function as part of our model fit, this criteria serves
to identify trim durations which minimize the presence of the
exponential ramp in our data.. We then re-run our previous
analysis in order to ensure that our aperture and bin sizes are
still optimal given this new trim duration.
2.2. Eclipse Model and Uncertainty Estimates
We generate our secondary eclipse light curves using the
routines from Mandel & Agol (2002), where we fix the planet-
star radius ratio, orbital inclination, eccentricity e, longitude
of periapse ω, and ratio of the orbital semi-major axis to the
stellar radius to their best-fit values from Lanotte et al. (2014).
We allow individual eclipse depths and center of eclipse times
FIG. 3.- Normalized Spitzer 3.6 and 4.5 µm light curves as a function
of time from the predicted center of eclipse, where we have divided out the
best-fit instrumental model shown in Fig. 1. The normalized flux is binned in
30 s (grey filled circles) and 5 minute (black filled circles) intervals, and best
fit eclipse model light curves are over plotted for comparison (solid lines).
to vary as free parameters in our fits to the 3.6 µm data. We
find that the eclipse depth in individual 4.5 µm observations
is consistent with zero, and therefore place a Gaussian prior
on the phase of the secondary eclipse in order to constrain the
best-fit eclipse time. We implement this prior as a penalty in
χ2 proportional to the deviation from the error-weighted mean
center-of-eclipse phase and corresponding uncertainty from
Knutson et al. (2011). Although we also calculate the best-fit
eclipse orbital phase using the e and ω values from Lanotte
et al. (2014) and find that it is consistent with the value from
Knutson et al. (2011), the corresponding uncertainty is sub-
stantially larger than that reported in Knutson et al. (2011).
This is not surprising, as the measured times of secondary
eclipse constrain ecos ω while esin ω is typically derived from
fits to radial velocity data and has larger uncertainties (e.g.
Pál et al. 2010; Knutson et al. 2014b). The uncertainties in
the values for e and ω reported in Lanotte et al. (2014) are
therefore likely to be dominated by the esin ω, while ecos ω
is well-measured from secondary eclipse photometry alone.
We chose not to include a prior for our 3.6 micron fits be-
cause we wanted to derive an independent estimate of the
eclipse center time and phase. However, to test the effect of
this choice, we repeat our 3.6 µm fits including this prior on
the eclipse phase and find that the measured eclipse depths
change by less than 0.1 sigma, as expected for cases where
the eclipse is detected at a statistically significant level.
We fit our combined eclipse and instrumental noise model
to each light curve using a Levenberg-Marquardt minimiza-
tion routine with uniform priors on all parameters except the
0.00010.00100.01000.00010.00100.01001101001000100000.00010.00100.0100110100100010000Bin SizeStandard Deviation of Residuals3.6 um, UT 2008 Jan 303.6 um, UT 2014 Jul 294.5 um, UT 2008 Feb 24.5 um, UT 2011 Jan 244.5 um, UT 2014 Aug 114.5 um, UT 2015 Feb 250.99951.00001.00051.00100.99951.00001.00051.0010-4-3-2-101230.99951.00001.00051.0010-4-3-2-10123Time From Predicted Eclipse Center [h]Relative Flux3.6 um, UT 2008 Jan 303.6 um, UT 2014 Jul 294.5 um, UT 2008 Feb 24.5 um, UT 2011 Jan 244.5 um, UT 2014 Aug 114.5 um, UT 2015 Feb 25GJ 436b
5
TABLE 2
BEST FIT ECLIPSE PARAMETERS
Fp/F∗,avg (ppm)a
Tbright (K)a
b
Ts
O −C (d)c
Fp/F∗ (ppm)
177± 31
133± 35
30± 36
37± 37
64± 45
1± 44
151± 27
29± 20
879+29
−27
< 633
λ (µm) UT Start Date
−0.0007± 0.0012
−0.0006± 0.0054
2008-01-30
2014-07-29
2008-02-02
2011-01-24
2014-08-11
2015-02-25
3.6
3.6
4.5
4.5
4.5
4.5
a We report the error-weighted mean eclipse depths at 3.6 and 4.5 µm. Brightness temperatures are calculated using a PHOENIX stellar model
interpolated to match the published stellar temperature and surface gravity from von Braun et al. (2012).
b BJDUTC - 2,450,000.
c Observed minus calculated eclipse times, where we have accounted for the uncertainties in both the measured and predicted eclipse times
as well as the XX s light travel time delay in the system. We calculate the predicted eclipse time using the best-fit eclipse orbital phase from
Knutson et al. (2011).
d We allow the eclipse times in this bandpass to vary as free parameters in our fit, but we use the orbital phase and corresponding uncertainty
from Knutson et al. (2011) as a prior constraint in the fit.
e 2σ upper limit based on the error-weighted average of the four 4.5 µm eclipse measurements.
4496.4888± 0.0012
6868.0655± 0.0054
4499.1334d
5585.7756d
6881.2856d
7079.5779d
4.5 µm eclipse time as described in the previous paragraph.
Our model includes nine pixel weight parameters, two eclipse
parameters, and a linear function of time in order to account
for long-term instrumental and stellar trends. We show the re-
sulting light curves and best-fit eclipse models after dividing
out the best-fit instrumental noise model and linear function
of time in Fig. 3. Uncertainties on model parameters are cal-
culated using a Markov chain Monte Carlo (MCMC) analysis
with 106 steps initialized at the location of the best-fit solu-
tion from our Levenberg-Marquardt minimization. We trim
any remaining burn-in at the start of the chain by checking
to see where the χ2 value of the chain first drops below the
median value over the entire chain, and trim all points prior to
this step. We find that in all cases our probability distributions
for the best-fit eclipse depths and times are Gaussian and do
not show any correlations with other model parameters. We
therefore take the symmetric 68% interval around the median
parameter value as our 1σ uncertainties.
3. ATMOSPHERIC MODELING
We use a combination of self-consistent modeling and re-
trieval algorithms to model the atmosphere of GJ 436b and
match its spectrum. The self-consistent modeling mirrors
that used in Morley et al. (2015); our suite of tools includes
a 1D radiative–convective model to calculate the pressure–
temperature structure, a photochemical model to calculate the
formation of hydrocarbons that may form hazes, and a cloud
model to calculate cloud mixing ratios, altitudes, and parti-
cle sizes. We calculate spectra in different geometries and
wavelengths using a transmission spectrum model, a thermal
emission spectrum model, and an albedo model. We also use
a retrieval model, CHIMERA (Line et al. 2012, 2013, 2014)
to explore the thermal emission spectrum. In the following
subsections we will briefly discuss each of these calculations.
We fit our models to the thermal emission and transmission
spectra separately and then analyze the regions of parameter
space where the same model parameters fit both the thermal
emission and transmission spectra.
3.1. 1D Radiative–Convective Model
We calculate the temperature structures of GJ 436b's at-
mosphere assuming radiative–convective equilibrium. These
models are more extensively described in McKay et al.
(1989); Marley et al. (1996); Burrows et al. (1997); Marley
et al. (1999, 2002); Fortney et al. (2005); Saumon & Marley
(2008); Fortney et al. (2008). Our opacity database for gases
is described in Freedman et al. (2008, 2014). We calculate
the effect of cloud opacity using Mie theory, assuming spher-
ical particles. Optical properties of sulfide and salt clouds and
soot haze are from a variety of sources and presented in Mor-
ley et al. (2012) and Morley et al. (2013).
To calculate P–T profiles for models with greater than 50×
solar metallicity, we make the same approximation as used
in Morley et al. (2015). We multiply the total molecular gas
opacity by a constant factor (e.g. we multiply the 50× so-
lar opacities by 6 to approximate the opacity in a 300× solar
composition atmosphere). We change the abundances of hy-
drogen and helium separately to calculate collision-induced
absorption. This approximation is appropriate for the results
explored here; for future work, e.g. comparing models to
JWST data, new k-coefficients at 100–1000OE solar metallic-
ity should be used.
3.2. Equilibrium Chemistry
After calculating the pressure–temperature profiles of mod-
els with greater than 50× solar metallicity, we calculate the
gas abundances assuming chemical equilibrium along that
profile. We use the Chemical Equilibrium with Applications
model (CEA, Gordon & McBride 1994) to compute the ther-
mochemical equilibrium molecular mixing ratios (with appli-
cations to exoplanets see, Visscher et al. (2010); Line et al.
(2010); Moses et al. (2011); Line et al. (2011); Line & Yung
(2013)). CEA minimizes the Gibbs Free Energy with an el-
emental mass balance constraint given a local temperature,
pressure, and elemental abundances. We include molecules
containing H, C, O, N, S, P, He, Fe, Ti, V, Na, and K. We
account for the depletion of oxygen due to enstatite conden-
sation by removing 3.28 oxygen atoms per Si atom (Burrows
& Sharp 1999). When adjusting the metallicity all elemental
abundances are rescaled equally relative to H, ensuring that
the elemental abundances sum to one.
3.3. Photochemical Haze Model
We use results from photochemical modeling in Line et al.
(2011). Briefly, the computations use the Caltech/JPL pho-
tochemical and kinetics model, KINETICS (a fully implicit,
finite difference code), which solves the coupled continuity
equations for each species and includes transport via both
molecular and eddy diffusion (Allen et al. 1981; Yung et al.
6
1984; Moses et al. 2005). We use results for 50× solar com-
position, Kzz=108 cm2/s (Figures 5, 6 and 7 in Line et al.
(2011)).
We follow the approach developed in Morley et al. (2013)
and used for GJ 1214b in Morley et al. (2015) to calculate the
locations of soot particles based on the photochemistry. We
sum the number densities of the five soot precursors (C2H2,
C2H4, C2H6, C4H2, and HCN) to find the total mass in soot
precursors. We assume that the soots form at the same alti-
tudes as the soot precursors exist: we multiply the precursors'
masses by our parameter fhaze (the mass fraction of precursors
that form soots) to find the total mass of the haze particles in
a given layer. We vary both fhaze and the mode particle size
as free parameters, and calculate the optical properties of the
haze using Mie theory.
3.4. Sulfide/Salt Cloud Model
To model sulfide and salt clouds, we use a modified version
of the Ackerman & Marley (2001) cloud model (Morley et al.
2012, 2013, 2015). Cloud material in excess of the saturation
vapor pressure of the limiting gas is assumed to condense into
spherical, homogeneous cloud particles. We extrapolate the
saturation vapor pressure equations from Morley et al. (2012)
to high metallicites, which introduces some uncertainties but
serves as a reasonable first-order approximation for the forma-
tion of these cloud species. Cloud particle sizes and vertical
distributions are calculated by balancing transport by advec-
tion with particle settling.
3.5. Thermal Emission Spectra
We use a radiative transfer model developed in Morley
et al. (2015) to calculate the thermal emission of a planet
with arbitrary composition and clouds. Briefly, this model
includes the C version of the open-source radiative transfer
code disort (Stamnes et al. 1988; Buras et al. 2011) which
uses the discrete-ordinate method to calculate intensities and
fluxes in multiple-scattering and emitting layered media.
3.6. Albedo Spectra
We calculate albedo spectra following the methods de-
scribed in Toon et al. (1977, 1989); McKay et al. (1989);
Marley et al. (1999); Marley & McKay (1999); Cahoy et al.
(2010). Here, we use the term geometric albedo to refer to
the albedo spectrum at full phase (α=0, where the phase an-
gle α is the angle between the incident ray from the star to the
planet and the line of sight to the observer):
Ag(λ) =
Fp(λ, α = 0)
F(cid:12),L(λ)
(2)
where λ is the wavelength, Fp(λ, α = 0) is the reflected flux at
full phase, and F(cid:12),L(λ) is the flux from a perfect Lambert disk
of the same radius under the same incident flux.
3.7. Retrieval Model
To more thoroughly explore the chemically plausible pa-
rameter space allowed by the emission spectrum, we employ
the chemically consistent atmospheric retrieval scheme de-
scribed in Kreidberg et al. (2015) and Greene et al. (2016)
based on the CHIMERA (Line et al. 2013, 2014) emission
forward model. The retrieval uses the 6 parameter analytic
radiative equilibrium temperature profile scheme of Parmen-
tier & Guillot (2014) (see Line et al. (2013) for implemen-
tation within the emission retrieval) where the free parame-
ters are the infrared opacity (κIR), the ratio of the visible to
FIG. 4.- Eclipse depths in the 6 Spitzer bandpasses from the literature and
this work. Different publications are offset slightly in wavelength for clarity;
darker colors indicate later years.
TABLE 3
UNIFORM PRIOR RANGES ON THE
RETRIEVED PARAMETERS
parameter
range
log(κIR)[cm2/g]
log(γ1, γ2)
α
β
Tint (K)
M/H
log(C/O)a
log(Pquench) [bar]
−3 to 0
−3 to 2
0 to 1
0 to 2
100 to 400
10−4 to 104× solar
−2 to 2
−6 to 1.5
a Solar log(C/O) is −0.26.
infrared opacity for two visible streams (γ1, γ2), the parti-
tioning between the two visible streams (α), scaling to the
top-of-atmosphere irradiation temperature (β, to accommo-
date for the unknown albedo and redistribution), and finally
the internal temperature (Tint). These parameters are all free
parameters, not recalculated to be consistent with the derived
abundances (Line et al. 2012).
The molecular abundances are initially computed along the
temperature profile under the assumption of thermochemi-
cal equilibrium (using the Chemical Equilibrium with Appli-
cations routine, Gordon & McBride 1994;1996; Line et al.
(2010); Moses et al. (2011); Line et al. (2011)) given the bulk
atmospheric metallicity ([M/H]) and carbon-to-oxygen ratio
(C/O). To account for possible disequilibrium chemistry we
include a "quench pressure" parameter (Pquench) whereby the
abundances of H2O, CH4, and CO above the quench are fixed
at their quench pressure values, a valid representation of many
disequilibrium models (e.g., Moses et al. 2011; Line et al.
2011; Zahnle & Marley 2014). The temperature profile and
chemistry parameters result in a total of 9 free parameters.
Bayesian estimation is performed using a multi-modal nested
sampling algorithm (Feroz et al. 2009) implemented with the
PYMULTINEST routine (Buchner et al. 2014) recently em-
ployed in Line & Parmentier (2016), with generous uniform
priors on each parameter (see Table 3).
4. RESULTS
4.1. Observations
GJ 436b
7
red ∼11 assuming
an inadequate fit to the thermal emission (χ2
3 degrees of freedom). We show the thermal emission and
transmission spectra in Figure 6.
4.2.2. Equilibrium and disequilibrium chemistry
As has been discussed in the literature (Stevenson et al.
2010; Line et al. 2011; Moses et al. 2013), GJ 436b's high 3.6
µm flux and low 4.5 µm flux indicate that it likely has a high
abundance of CO and CO2 relative to CH4. Since equilib-
rium chemistry for an object at GJ 436b's temperature would
instead result in high abundances of CH4 at metallicities sim-
ilar to Neptune, this indicates that GJ 436b's chemistry is in
disequilibrium. This disequilibrium may be due to a combi-
nation of vertical mixing, photochemistry, and other effects
(Line et al. 2011). Here, we approximate the effect of dis-
equilibrium chemistry by 'quenching' the abundances of the
carbon species (CO, CO2, CH4) in the atmosphere at deep
pressures (10 bar), effectively setting the abundances of these
species to be constant through the atmosphere.
The resulting effect of disequilibrium chemistry on spec-
tra is shown in Figure 7. In equilibrium, the model predicts
that GJ 436b would be very faint at 3.6 µm, and progressively
brighter at redder wavelengths. In disequilibrium, as is ob-
served in the data, the planet is predicted to be brighter at 3.6
µm due to decreased absorption by CH4. In general, even the
models that include disequilibrium chemistry overpredict the
brightness at 4.5 µm compared to the observed flux, despite
the higher abundance of CO and CO2 in disequilibrium.
4.2.3. Metallicity
Increasing the metallicity of GJ 436b's atmosphere allows
us to fit both the thermal emission and transmission spectrum
more accurately. There are two reasons for this. As has been
discussed at length in Moses et al. (2013), high metallicity
atmospheres are predicted, in equilibrium or disequilibrium,
to have higher abundances of CO and CO2 relative to CH4.
Pushing the chemistry to CO/CO2-rich compositions is cru-
cial to match GJ 436b's thermal emission. We show this ef-
fect in Figure 8; models at high metallicities have higher flux
at 3.6 and 8 µm due to the change in chemistry. We find that
this effect partially saturates at metallicities greater than 300×
solar.
High metallicities also make it much easier to flatten the
transmission spectrum of GJ 436b sufficiently to match the
featureless HST/WFC3 transmission spectrum even in the ab-
sence of clouds (Knutson et al. 2014a). In Figure 9 we show
cloud-free models for different metallicities. While at metal-
licities lower than 1000× solar metallicity clouds are required
to sufficiently flatten the spectrum, for models above 1000×
solar metallicity even cloud-free models have high enough
mean molecular weights that the size of the features, which
scale according to the scale height of the atmosphere, are
small enough that they appear featureless at the S/N of the
data.
4.2.4. Tidal heating
As a Neptune-sized planet orbiting an old star, without an
additional energy source, GJ 436b's interior temperature Tint
would be ∼60 K, slightly warmer than Neptune which has
a Tint∼50 K (Fortney et al. 2007). However, GJ 436b is on
an eccentric orbit (e∼0.15) despite orbiting its star at a semi
major axis where it is predicted to have a tidally circularized
orbit, indicating that its interior may still be heated by tidal
FIG. 5.- Pressure–temperature profiles with condensation curves. All
models are cloud-free with 300× solar composition. Solid lines show models
with Tint=100, 240, and 400 K and planet-wide heat redistribution. Dash-dot
lines show models with the same Tints but with no heat redistribution (day-
side temperature). Condensation curves show where the vapor pressure of a
gas is equal to the saturation vapor pressure; cloud material condenses where
the P–T profile intersections a condensation curve.
The new eclipse depths are shown in Table 2 and Figure
4. Our eclipse depths of 151±27 ppm at 3.6 µm and 29+20
−16
ppm at 4.5 µm are consistent to 1-σ with those published in
Lanotte et al. (2014) (177±45 and 28+25
−18 ppm respectively),
with a moderate reduction in the uncertainties in both bands.
This result serves as confirmation of the high flux at 3.6 µm
compared to 4.5 µm.
4.2. Self-Consistent Modeling
We ran a variety of models from 50–1000× solar metallic-
ity, varied heat redistribution (planet-wide average and day-
side average), internal temperatures (Tint) from 100–400 K,
with clouds ( fsed=0.01–1), and hazes with particle sizes from
0.01–1 µm and fhaze from 1–30%. We compare each model
to the thermal emission photometry from this work (3.6 and
4.5 µm) and from Lanotte et al. (2014) (5.6, 8.0, 16 µm), us-
ing a chi-squared analysis to assess relative goodness-of-fit
between the models.
We show example pressure–temperature profiles along with
cloud condensation curves in Figure 5. Raising the internal
temperature, Tint, increases the temperature of the deep at-
mosphere (P(cid:38)0.1 bar). The heat redistribution of incident
stellar flux controls the temperature in the upper atmosphere.
GJ 436b's profile crosses condensation curves of sulfides and
salts, suggesting that if the atmosphere is cloudy, those clouds
may be composed of Na2S, KCl, and ZnS.
4.2.1. Best-fit fiducial model
Of the 288 models in our grid of cloudy and cloud-free
planets, our nominal best-fit set of parameters are:
• 1000× solar metallicity
• Tint=240 K
• fsed=0.3 sulfide/salt clouds
• disequilibrium chemistry via quenching
• full heat redistribution (planet-wide average PT profile)
This model provides an excellent fit to the transmission
red <1 assuming 3 degrees of freedom), though
spectrum (χ2
2004006008001000120014001600temperature (K)6543210123log(pressure) (bar)MnSNa2SZnSCrKClH2OTint=400 K240 K100 Kplanet-wide redistributionno redistribution8
FIG. 6.- Best-fit thermal emission and transmission spectra. Top panel: Thermal emission spectrum of the best-fit model from the suite of forward models
compared to the data. The model is shown as a green line, with synthetic model photometry shown as horizontal lines at the central wavelength of the filter. Data
are shown as black points with 1-σ error bars. The filter functions for the photometry are shown as gray lines in the top panel. Bottom panels: Transmission
spectrum of the same best-fit thermal emission model from the suite of forward models compared to the data. The model is shown as a green line in both panels.
The HST/WFC3 transmission spectrum is shown as black points with 1-σ error bars in the bottom panel.
dissipation. Moses et al. (2013) and Agúndez et al. (2014)
both considered the effect of tidal heating, noting that a hotter
interior changes the chemistry of the deep interior and there-
fore the resulting emission spectrum.
Increasing Tint tends to move the deep P–T profile (see Fig-
ure 5) to regions with high CO/CO2 and lower CH4 abun-
dances, which allows us to better match the observed spec-
trum. Heating the deep atmosphere also increases the effec-
tive temperature of the atmosphere by changing the P–T pro-
file, increasing flux at all Spitzer wavelengths. This effect is
shown in Figure 11 for three different Tint values (100, 240,
and 400 K). Best-fit models cluster around Tint=240 K, a tem-
perature that allows us to match the 3.6, 5.6, and 8.0 µm points
relatively well, while over predicting the 4.5 µm flux some-
what.
We note that this is the first indication that the internal tem-
perature of a planet has an important and observable effect on
the emission spectrum of a transiting planet.
4.2.5. Clouds
Clouds increase opacity across all wavelengths as (rela-
tively) gray absorbers. This means that including clouds de-
creases flux between absorption features (e.g. at 3.6 and 8.0
µm for GJ 436b's composition) and somewhat less signifi-
cantly at the locations of absorption features where the planet
is already dark. Thinner clouds ( fsed=0.3–1 in our param-
eterization) alter the spectrum slightly, while thicker clouds
( fsed≤0.1) create a blackbody-like spectrum with the temper-
ature of the top of the cloud. Comparing this to the observed
photometry of GJ 436b, these thick clouds significantly under
predict the flux at 3.6 µm especially.
In transmission, clouds flatten the spectrum without in-
creasing the mean molecular weight of molecular gas in the
atmosphere. As discussed above, for metallicities ∼1000×
solar, no additional cloud opacity is needed to match the fea-
red∼1 for all models). At 300× solar
tureless spectrum (χ2
metallicity, thin clouds ( fsed=1) adequately obscure the spec-
tral features, whereas for a Neptune-like 100× solar composi-
tion, fsed=0.3 clouds are required. In the Ackerman & Marley
(2001) prescription, lower fsed values indicate less efficient
23456789101520wavelength (¹m)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]1000£, Tint=240 K, fsed=0.3, quenched0.51251020wavelength (µm)4.2554.2604.2654.2704.2754.2804.2854.290radius (R⊕)1000×, Tint=240 K, fsed=0.3, quenched1.151.251.351.451.551.65wavelength (µm)20015010050050100150relative transit depth, ppmGJ 436b
9
FIG. 7.- Effect of chemistry on thermal emission spectrum. Both mod-
els assume 300× solar metallicity,
fsed=1 sulfide/salt clouds, planet-wide
heat redistribution, and Tint=240 K. The darker blue line and horizontal bars
show a model spectrum and photometry assuming equilibrium chemistry; the
lighter blue line and horizontal bars show the same model, but with the chem-
istry quenched at the 10 bar abundances throughout the atmosphere.
FIG. 9.- Effect of metallicity on transmission spectrum. Each model is
cloud-free, with planet-wide heat redistribution, equilibrium chemistry, and
Tint=240 K. Metallicities of 100, 200, 300, and 1000× solar metallicity are
shown. Increasing the metallicity decreases the CH4 abundance and increases
CO and CO2 abundance.
FIG. 8.- Effect of metallicity on thermal emission. Each model assumes
equilibrium chemistry, fsed=1 sulfide/salt clouds, planet-wide heat redistribu-
tion, and Tint=240 K. Metallicities of 100, 300, and 1000× solar metallicity
are shown. Increasing the metallicity decreases the CH4 abundance and in-
creases CO and CO2 abundance.
sedimentation, causing smaller particles sizes and more lofted
clouds.
4.2.6. Photochemical Hazes in GJ 436b
We investigate the effect of photochemical hazes on the
thermal emission spectrum of GJ 436b. Morley et al. (2015)
showed that it is possible for optically thick photochemical
hazes (such as those postulated to exist in GJ 1214b) to cause
a temperature inversion in the upper atmospheres of plan-
ets. This can change the spectrum such that molecules that
would normally be seen in absorption in a planet without a
temperature inversion such as methane are actually seen in
emission in an atmosphere with a temperature inversion. We
tested whether this process could be happening on GJ 436b
and causing the observed thermal emission.
The results of this investigation are summarized in Figure
13. The top panel shows the thermal emission of the planet
alone. We find that it is possible to create a temperature inver-
sion with dark soot-like photochemical haze in GJ 436b, espe-
cially for relatively small particle sizes. As expected, methane
is seen in emission, significantly brightening the model spec-
trum at 3.6 µm compared to a haze-free model. As in Morley
FIG. 10.- Abundances of major carbon-bearing species in chemical equi-
librium. All models have a composition of 1000× solar metallicity and a
planet-wide average PT profile. Different Tint values are shown with differ-
ent line styles, and each molecule (CH4, CO, CO2) is shown in a different
color. The fiducial quench pressure used in the self-consistent modeling is
shown as a horizontal dashed line. Note that increasing the internal tempera-
ture decreases the CH4 abundance in the deep atmosphere.
et al. (2015), CO2 at 4.3 µm is also predicted to be seen in
emission at Neptune-like metallicities (in this case 50× solar
metallicity). In the bottom panel, we show the planet-star flux
ratio; here it becomes clear that the hazy model does not fit
the observations significantly better than the haze-free model.
In particular, the model spectrum is much fainter than the
planet's 3.6 µm photometric point. The 4.5 µm flux, despite
the significant changes to the shape of the spectrum across the
bandpass, remains nearly identical across the range of hazy
models tested.
In general, we find that even though a temperature inversion
in a methane-rich atmosphere can increase the 3.6 µm flux, it
is not a significant enough effect to match the observed flux,
and, furthermore, the flux within the 4.5 µm region can also
increase due to emission in the CO2 bandpass. We conclude
that photochemical hazes cannot erase the need for an atmo-
23456789101520wavelength (¹m)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]quenched chemistryequilibrium chemistry23456789101520wavelength (¹m)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]100£ solar300£ solar1000£ solar1.151.251.351.451.551.65wavelength (µm)2001000100200relative transit depth, ppm100×, χred=5.4300×, χred=2.01000×, χred=0.910-810-710-610-510-410-310-210-1100number mixing ratio10-610-510-410-310-210-1100101102103pressure (bar)quench pressureCH4, Tint=100 KCO2, Tint=100 KCO, Tint=100 KCH4, Tint=240 KCO2, Tint=240 KCO, Tint=240 KCH4, Tint=400 KCO2, Tint=400 KCO, Tint=400 K10
FIG. 11.- Effect of tidal heating on thermal emission. Each model as-
sumes 300× solar metallicity, quenched chemistry, fsed=1 sulfide/salt clouds,
and planet-wide heat redistribution. The tidally heated atmospheres (240 and
400 K) have higher abundances of CO and CO2 and lower abundances of
CH4 due to the hotter deep atmosphere (where the chemistry is quenched).
Tidal heating also increases the Teff of the planet by changing the temperature
profile, increasing the emergent flux at all wavelengths.
FIG. 12.- Effect of sulfide/salt clouds on thermal emission. Each model
uses the same pressure-temperature profile and assumes 300× solar metallic-
ity, quenched chemistry, planet-wide heat redistribution, and Tint=240 K. A
cloud-free model and cloudy models with fsed=0.03 to 1 are shown. Cloud
opacity decreases the thermal emission across the spectrum. Models with
moderate clouds ( fsed=0.3 to 1) fit the Spitzer points best.
sphere with significant CO and CO2 and a low abundance of
CH4. This required low-CH4 atmospheric composition, in
turn, reduces the likelihood that carbon-based photochemi-
cal hazes will be significant in the atmosphere (Fortney et al.
2013).
4.3. Retrievals
We have shown in Section 4.2 that we favor models at high
metallicity, with both disequilibrium chemistry and tidal heat-
ing; these three properties combine to maximize the CO/CO2
abundances and minimize CH4 abundance, allowing the mod-
els to match approximately with the measured photometry.
Retrieval models provide a quantitative way to test these con-
clusions and fully explore parameter space beyond our self-
consistent model grids.
We find that retrieval methods draw similar conclusions to
the self-consistent modeling; GJ 436b appears to be very high
metallicity, with evidence for both deeply-quenched disequi-
librium chemistry and thermal heating of the deep interior.
For the dayside thermal emission spectrum, the best-fit re-
FIG. 13.- Effect of photochemical hazes on thermal emission. The top
panel shows the emergent flux from the planet. All models have 50× solar
metallicity, equilibrium chemistry, and planet-wide heat redistribution. The
gray line shows a cloud-free model, and the colored lines show a progression
of hazy models with hazy-forming efficiency parameter fhaze varying from 1
to 30%. The bottom panel shows the same models, but dividing by the flux
of the host star to compare to the measured photometry.
trieved solution has a goodness-of-fit divided by number of
data points χ2/N=2.02, compared to χ2/N=4.54 for the best
self-consistent thermal emission spectrum, indicating a sig-
nificantly improved fit. A comparison of the Bayesian Infor-
mation Criteria (BIC) reveals that that additional parameters
used in the retrieval fitting are necessary, with a BIC of ∼26
for the best-fit retrieved solution and ∼34 for the best-fit self-
consistent solution.
4.3.1. Retrieved Posterior Probability Distributions
Retrieved posterior probability distributions and correla-
tions are shown in the stair-pair plot in Figure 14 for 5 of the
9 free parameters in the retrieval: β, Tint, [M/H], log(C/O),
log(Pquench). The best-fit models have:
• High metallicity. The maximum likelihood model has
a metallicity of ∼6000× solar metallicity, with a 3-σ
lower limit on the metallicity of 106× solar.
• Disequilibrium chemistry. The maximum likelihood
model has a quench pressure around 9 bar (with a wide
range of values for Pquench allowed).
• Enhanced internal temperature. The maximum likeli-
hood Tint is 336 K (with large uncertainties), indicating
23456789101520wavelength (¹m)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]Tint=100 KTint=240 KTint=400 K23456789101520wavelength (¹m)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]fsed=0.1fsed=0.3fsed=1cloud-free123456789101520wavelength (µm)05001000150020002500flux (W/m2/µm)50×, haze-free0.03 µm, fhaze=1%0.03 µm, fhaze=3%0.03 µm, fhaze=10%0.03 µm, fhaze=30%23456789101520wavelength (µm)0.00000.00020.00040.00060.00080.00100.00120.00140.00160.0018Fp/Fs[3.6][4.5][5.8][8.0][16]50×, haze-free0.03 µm, fhaze=1%0.03 µm, fhaze=3%0.03 µm, fhaze=10%0.03 µm, fhaze=30%GJ 436b
11
FIG. 14.- Posterior probability distributions and correlations. The top panel (histogram) shows the posterior probability distribution for each parameter,
marginalized over all other parameters. The other panels show 2D contour plots that represent the correlations between each pair of parameters, where the
regions from darkest to lightest represent the 1-, 2-, and 3-σ contours.
that tidal heating may be increasing GJ 436b's internal
temperature, in agreement with the tidally heated self-
consistent models.
• Solar C/O ratio. The maximum likelihood C/O ratio
is 0.70, with a sharp cut-off at higher C/O ratios and a
long tail to lower C/O ratios.
In Figure 15 we compare the retrieved P–T profile to self-
consistent models at 300 × solar metallicity. We find that re-
trieved profile is in remarkable agreement with self-consistent
models that include the effect of tidal heating in the deep in-
terior. Our best-fit Tint from the self-consistent modeling ap-
proach (240 K) falls within the 2-σ range of the retrieved pro-
file.
The contribution functions for each of the Spitzer band-
passes are also shown in Figure 15. 3.6 µm probes the deepest
pressures, probing pressures as high as 1 bar. As expected,
comparing the contribution functions to the range of P–T pro-
files found by the retrieval, the spread in allowed P–T profiles
increases for pressures deeper than 1 bar. The other wave-
lengths probe lower pressures of the atmosphere, with 5.8 and
8.0 µm centered around 0.05 bar and 16 µm centered around
0.003 bar. The 4.5 µm bandpass has the largest range of pres-
sures, with a peak at deep pressures (0.2 bar) and a long tail
to low pressures, unsurprising given that the band covers the
spectrum where the modulation is the greatest.
Figure 16 shows the best-fit retrieved range of spectra com-
pared to both the data and the best-fit self-consistent model.
The retrieved best-fit is statistically and by-eye a somewhat
better fit to the data than the self-consistent models. In par-
ticular, it has higher flux at 3.6 µm and lower flux at 4.5 µm.
Both the retrieved and self-consistent models fit the 5.6 and
8.0 µm points well; the 16 µm photometry is underestimated
by both models, though the error bar is large.
5. DISCUSSION
5.1. Predictions for Reflected Light Spectra
Cloud properties have the strongest effect on the predicted
reflected light spectrum of GJ 436b. Cloud-free models are
dark from 0.6–1µm (Ag<1%) and somewhat brighter (up to
Ag ∼10%) at bluer wavelengths, as is generally true for cloud-
less giant planets (Marley et al. 1999; Sudarsky et al. 2000).
Thinner clouds ( fsed=0.3–1) are brighter with albedos be-
tween a few percent and tens of percent. Thicker clouds
( fsed=0.1) have the brightest albedos from 0.6 to 1µm, up to
nearly 30%. Some example cloudy spectra are shown in the
top panel of Figure 17.
Other properties have weaker effects on the reflected light
spectrum for this planet. For example, models with metal-
licities from 100–300× solar metallicity are shown in the bot-
tom panel of Figure 17. Increasing the metallicity (which also
changes the cloud) increases the geometric albedo across the
spectrum.
5.2. Are Very High Metallicities Reasonable?
We find that the best-fit atmospheric models have high
metallicities, but it remains to be seen whether these values
12
FIG. 15.- Pressure–temperature profiles and contribution functions for each bandpass. The left panel shows pressure-temperature profiles of both retrieved
and self-consistent models. The black line indicates the median retrieved profile while the dark and light gray shaded regions represent the 1- and 2-σ confidence
regions respectively. The colored lines show self-consistent models with planet-wide heat redistribution and Tint of 100, 240, and 400 K. Note the good
agreement between the tidally heated (240–400 K) models and the retrieved profile. The right panel shows contribution functions for each of the five bandpasses
for a representative retrieval model. The shortest wavelength 3.6 µm band probes the deepest wavelengths while the 16 µm band proves the shallowest.
FIG. 16.- Retrieved data compared to data and best-fit self-consistent model. The pink line and shaded dark and light pink regions are the median fit, 1-σ, and
2-σ confidence intervals respectively. The green line is the best-fit self-consistent model (300× solar metallicity, Tint=240 K, fsed=0.3, quenched disequilibrium
chemistry).
are physically realistic. GJ 436b has a different host star, equi-
librium temperature, and orbit than the ice giants in our own
solar system, so it likely formed and evolved in very different
conditions. The maximum metal-enrichment of the envelope
of a Neptune-mass exoplanet is not yet known. Studies of this
to date, including Fortney et al. (2013), have suggested that
a diverse range of outcomes might be expected for planets in
this intermediate mass regime between Earth and Saturn, with
potentially high atmospheric enrichments in some cases.
Furthermore, because of the uncertainty in the internal en-
tropy of GJ 436b, its mass and radius do not provide strong
limits on the metal-enrichment of the envelope. Nettelmann
et al. (2010) find that a minimum H/He fraction of 10−3 Mp
is necessary to match the radius. This very low H/He fraction
would require a warm planetary interior, as is favored by the
best-fit thermal emission spectra in this work.
In Figure 18 we show the results from interior models
(Thorngren et al. 2015), which show the envelope H/He mass
fraction required to match GJ 436b's mass and radius mea-
surements for core masses from 0 to 20 M⊕. These one-
dimensional interior models include an inert core composed
of either 50% rock and 50% ice or 100% water ice. Each
model has a homogenous convective envelope made of a
H/He-rock-ice or H/He-ice mixture. We use equations of state
from Thompson (1990) (water and water-rock) and Saumon
et al. (1995) (H/He), and atmospheric models from Fortney
et al. (2007). Neither the pure water or water-rock equations
of state are dependent on temperature; since GJ 436b has a hot
interior, we expect that the true equation of state of its interior
4006008001000120014001600temperature (K)654321012log(pressure) (bar)Tint=400 K240 K100 Kmedian1-σ2-σ0.00.20.40.60.81.0contribution function654321012log(pressure) (bar)[3.6][4.5][5.8][8.0][16]23456789101520wavelength (µm)0.00000.00020.00040.00060.00080.00100.00120.00140.0016Fp/Fsbest-fit self-consistent modelmedian retrieved spectrum1-σ2-σGJ 436b
13
velope metallicity is somewhat lower than 1000× solar metal-
licity. For a 20 M⊕ core, the minimum envelope metallicity
required to match the observed mass and radius is ∼300× so-
lar metallicity (though the envelope could be less enriched if
the core is somewhat larger than 20 M⊕. These calculations
suggest that the very high metallicities above 1000× solar
metallicity explored by our retrieval models are not physically
realistic for this planet, but metallicities above 300× solar are
favored.
Very high metallicities are only possible if accretion and
subsequent enrichment is dominated by rocky rather than icy
materials; Fortney et al. (2013) show that if the majority of ac-
cretion is from icy material, the hydrogen in those ices is also
accreted and the maximum metal-enrichment is about ∼600×
solar metallicity. Though we cannot currently distinguish be-
tween compositions less than or greater than 600× solar com-
position, if GJ 436b is indeed very metal-enhanced, it likely
formed in a region with more refractory than volatile materi-
als available.
5.3. Role of JWST Spectral Observations
JWST will amplify our understanding of warm Neptunes
like GJ 436b by providing spectra instead of photometry,
breaking some of the current degeneracies. For example,
examining the spectra in Figure 16, it is clear that models
with very different spectra can have very similar photometry.
JWST may also allow us to detect molecules that are not cur-
rently included in most models; for example, Shabram et al.
(2011) showed that if species such as C2H2 and HCN exist in
the atmosphere of GJ 436b, their abundances could be con-
strained by measuring the widths of features at 1.5, 3.3, and 7
µm.
Greene et al. (2016) quantify our ability to constrain planet
properties of a wider variety of atmospheres including hot
Jupiters, warm Neptunes, warm sub-Neptunes, and cool su-
per Earths with JWST and find that the mixing ratios of major
species in warm Neptunes like GJ 436b can be constrained to
within better than 1 dex with a single secondary eclipse ob-
servation for each wavelength region from 1–11 µm.
5.4. Measuring Internal Dissipation Factor Using Tint
Measuring Tint of GJ 436b using atmospheric models al-
lows us to approximate the dissipation factor in GJ 436b's
interior, Q(cid:48). Q(cid:48) is defined as 3Q/2k2, where Q is the qual-
ity factor and k2 is the Love number of degree 2 (Goldreich
& Soter 1966). Our best-fit Tint from the retrieval analysis
is 336 K. Agúndez et al. (2014) calculated relations between
Tint and Q(cid:48) assuming obliquities of 0 and 15 degrees and 3
different rotation speeds (1:1 resonance, 3:2 resonance, and
pseudo-synchronous). Assuming Tint∼300–350 K, their cal-
culations suggest that Q(cid:48) ∼ 2 × 105–106. These values are
somewhat larger than the value of Q(cid:48) that has been measured
using Neptune's satellites of between 3.3×104 and 1.35×105
(Zhang & Hamilton 2008).
If this high value for Q(cid:48) is correct, this has significant im-
plications for both the structure of GJ 436b itself and the evo-
lution of the planetary system. In particular, a high Q(cid:48) is con-
sistent with a tidal circularization timescale that is longer than
the age of the system, allowing GJ 436b to maintain its non-
zero eccentricity (e ∼0.15) without invoking another object in
the system (Jackson et al. 2008; Batygin et al. 2009). This is
consistent with the observations to date that have not found
a third body in the GJ 436 system despite extensive searches
FIG. 17.- Predicted albedo spectra. Top panel shows models with
300×solar metallicity, Tint=240 K. A cloud-free model and models with
cloud parameter fsed from 0.03 to 1 are shown. Bottom panel shows mod-
els with Tint=240 K, fsed=0.3. Metallicities from 100 to 300 × solar are
shown.
FIG. 18.- Core mass vs. envelope H/He fraction required to match GJ
436b's measured mass and radius. Models that include a rock/ice core and
mix of H/He and rock/ice in the envelope are shown as solid lines, while
models that include a water ice core and mix of H/He and water in the enve-
lope are shown as dashed lines. Different colors represent different interior
temperature Tint. Approximate conversions between envelope H/He fraction
and atmospheric metallicity for 1000 and 300× solar metallicity are shown
at the right.
will be more accurately matched by the water ice equation of
state, even though the composition is likely a mix of heavy
elements.
We find that for the best-fit Tint of 336 K the maximum en-
0.30.40.50.60.70.80.91.0wavelength (µm)0.00.10.20.30.40.5geometric albedo300× solar, Tint=240 Kcloud-freefsed=1fsed=0.3fsed=0.10.30.40.50.60.70.80.91.0wavelength (µm)0.00.10.20.30.40.5geometric albedoTint=240 K, fsed=0.3100 × solar200 × solar300 × solar05101520CoreMass(EarthMasses)0.050.100.150.200.250.300.350.40EnvelopeH/HeFraction100K240K336K400K50%Rock,50%IceWater~1000x ~300x 14
If GJ 436b's Q does differ significantly
(see Section 1.3).
from Neptune's, this potentially implies a structural difference
between hot Neptunes on short orbits and the ice giants in our
own solar system.
5.5. Condensation of graphite
As has been discussed in, e.g., Moses et al. (2013), cool
high metallicity atmospheres may have regions that are stable
for the condensation of graphite. Indeed, the very high metal-
licity models favored by the retrieval models do indeed cross
the graphite stability curve above 0.1 bar. While the effect of
this condensation is beyond the scope of this work, the major
effects would be twofold. First, the graphite condensation will
deplete the carbon reservoir, decreasing the CO abundance in
the upper atmosphere.
In addition, the condensed graphite
may form into cloud particles with their own opacity. Like
other clouds and hazes, graphite clouds would likely decrease
the size of features in transmission spectra and thermal emis-
sion spectra, and may either increase or decrease the albedo
depending on the optical properties of the graphite particles.
5.6. Spatially Inhomogeneous Clouds
Planetary atmospheres are by their nature three dimensional
and complex, and clouds in these atmospheres may be non
uniformly located. In particular, the terminators of a planet
have different circulation patterns and temperatures than the
substellar point (e.g., Lewis et al. 2010; Kataria et al. 2016).
It has recently been shown that non-uniform clouds on the
terminators of planets will affect interpretation of transmis-
sion spectra (Line & Parmentier 2016). Furthermore, fitting
the same one-dimensional model to both the thermal emission
and transmission spectra may not be an accurate assumption.
For example, the terminators may be cloudy while the ther-
mal emission is dominated by a relatively cloud-free region.
This effect should be investigated in the future, especially
with higher resolution and signal-to-noise thermal emission
and transmission spectra from JWST.
6. CONCLUSION
We have presented new observations of GJ 436b's thermal
emission at 3.6 and 4.5 µm, which are in agreement with pre-
vious analyses from Lanotte et al. (2014) and reduce the un-
certainties of GJ 436b's flux at those wavelengths. For the first
time, we combine these revised data with Spitzer photometry
from 5.6 to 16 µm and transmission spectra from HST/WFC3
and compare these data to both self-consistent and retrieval
models. We vary the metallicity, internal temperature from
tidal heating, disequilibrium chemistry, heat redistribution,
and cloud properties. We find that our nominal best-fitting
self-consistent model has 1000× solar metallicity, Tint=240
K, fsed=0.3 sulfide/salt clouds, disequilibrium chemistry, and
planet-wide average temperature profile, but this model does
not provide an accurate fit to the observations. Retrieval mod-
els find a statistically better fit to the ensemble data than the
self-consistent model, with parameters in general agreement
with the self-consistent approach: all signs point to a high
metallicity, with best fits above several hundred times solar
metallicity, and tidal heating warming its interior, with best-
fit Tint∼300–350 K. These results are consistent with results
from interior models to match the mass and radius, with core
masses around 10 M⊕.
While Neptune has been measured based on its methane
abundance to have an atmospheric carbon enhancement of
∼100× solar, repeated observations of both the thermal emis-
sion and transmission spectra of the first exo-Neptune to be
studied in detail, GJ 436b, have demonstrated that it likely
has a significantly higher metallicity. Neptune itself may ac-
tually be more enhanced in other elements than it is in car-
bon; Luszcz-Cook & de Pater (2013) infer a 400–600× solar
enhancement in oxygen from microwave observations of up-
welled CO in Neptune, though this cannot be verified with
infrared spectra since oxygen is frozen into clouds. Studies of
warmer exoplanet atmospheres will allow us to spectroscopi-
cally measure abundances of these molecules like oxygen that
are locked into clouds in the cold ice giants of our solar sys-
tem, potentially revealing unexpected patterns in the metal-
enrichments of these intermediate-mass objects.
An interesting new paradigm for this class of intermediate-
sized planet is now being pieced together: we suggest that
Neptune-mass planets may be more compositionally diverse
than previously imagined. High quality data across of range
of Neptune-mass planets with different temperatures and host
stars will be critical to investigate the diversity of this class of
planets.
We thank the anonymous referee for their helpful sugges-
tions. We also thank Konstantin Batygin and Greg Laughlin
for helpful discussions that improved the paper. This work
was performed in part under contract with the Jet Propul-
sion Laboratory (JPL) funded by NASA through the Sagan
Fellowship Program executed by the NASA Exoplanet Sci-
ence Institute.
JJF acknowledges Hubble grants HST-GO-
13501.06-A and HST-GO-13665.004-A and NSF grant AST-
1312545. MSM acknowledges support of the NASA Origins
program. MRL acknowledges support provided by NASA
through Hubble Fellowship grant #51362 awarded by the
Space Telescope Science Institute, which is operated by the
Association of Universities for Research in Astronomy, In.,
for NASA, under the contract NAS 5-26555. This work is
based on observations made with the Spitzer Space Telescope,
which is operated by the Jet Propulsion Laboratory, California
Institute of Technology, under contract with NASA.
REFERENCES
Ackerman, A. S., & Marley, M. S. 2001, ApJ, 556, 872
Agúndez, M., Venot, O., Selsis, F., & Iro, N. 2014, ApJ, 781, 68
Allen, M., Yung, Y. L., & Waters, J. W. 1981, J. Geophys. Res., 86, 3617
Alonso, R., Barbieri, M., Rabus, M., et al. 2008, A&A, 487, L5
Ballard, S., Charbonneau, D., Deming, D., et al. 2010a, PASP, 122, 1341
Ballard, S., Christiansen, J. L., Charbonneau, D., et al. 2010b, ApJ, 716,
1047
Batygin, K., Laughlin, G., Meschiari, S., et al. 2009, ApJ, 699, 23
Beaulieu, J.-P., Tinetti, G., Kipping, D. M., et al. 2011, ApJ, 731, 16
Beust, H., Bonfils, X., Montagnier, G., Delfosse, X., & Forveille, T. 2012,
A&A, 545, A88
Buchner, J., Georgakakis, A., Nandra, K., et al. 2014, A&A, 564, A125
Buras, R., Dowling, T., & Emde, C. 2011, J. Quant. Spec. Radiat. Transf.,
112, 2028
Burrows, A., & Sharp, C. M. 1999, ApJ, 512, 843
Burrows, A., Marley, M., Hubbard, W. B., et al. 1997, ApJ, 491, 856
Butler, R. P., Vogt, S. S., Marcy, G. W., et al. 2004, ApJ, 617, 580
Cáceres, C., Ivanov, V. D., Minniti, D., et al. 2009, A&A, 507, 481
Cahoy, K. L., Marley, M. S., & Fortney, J. J. 2010, ApJ, 724, 189
Deming, D., Harrington, J., Laughlin, G., et al. 2007, ApJ, 667, L199
Deming, D., Knutson, H., Kammer, J., et al. 2015, ApJ, 805, 132
Demory, B.-O., Gillon, M., Barman, T., et al. 2007, A&A, 475, 1125
Feroz, F., Hobson, M. P., & Bridges, M. 2009, MNRAS, 398, 1601
Fortney, J. J., Marley, M. S., & Barnes, J. W. 2007, ApJ, 659, 1661
GJ 436b
15
Fortney, J. J., Marley, M. S., Lodders, K., Saumon, D., & Freedman, R.
Fortney, J. J., Marley, M. S., Saumon, D., & Lodders, K. 2008, ApJ, 683,
2005, ApJ, 627, L69
1104
Fortney, J. J., Mordasini, C., Nettelmann, N., et al. 2013, ApJ, 775, 80
Freedman, R. S., Lustig-Yaeger, J., Fortney, J. J., et al. 2014, ApJS, 214, 25
Freedman, R. S., Marley, M. S., & Lodders, K. 2008, ApJS, 174, 504
Gillon, M., Demory, B.-O., Barman, T., et al. 2007a, A&A, 471, L51
Gillon, M., Pont, F., Demory, B.-O., et al. 2007b, A&A, 472, L13
Goldreich, P., & Soter, S. 1966, Icarus, 5, 375
Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ, 817, 17
Jackson, B., Greenberg, R., & Barnes, R. 2008, ApJ, 681, 1631
Kammer, J. A., Knutson, H. A., Line, M. R., et al. 2015, ApJ, 810, 118
Kataria, T., Sing, D. K., Lewis, N. K., et al. 2016, ApJ, 821, 9
Knutson, H. A., Benneke, B., Deming, D., & Homeier, D. 2014a, Nature,
505, 66
Knutson, H. A., Madhusudhan, N., Cowan, N. B., et al. 2011, ApJ, 735, 27
Knutson, H. A., Fulton, B. J., Montet, B. T., et al. 2014b, ApJ, 785, 126
Kreidberg, L., Line, M. R., Bean, J. L., et al. 2015, ApJ, 814, 66
Lanotte, A. A., Gillon, M., Demory, B.-O., et al. 2014, A&A, 572, A73
Lewis, N. K., Showman, A. P., Fortney, J. J., et al. 2010, ApJ, 720, 344
Lewis, N. K., Knutson, H. A., Showman, A. P., et al. 2013, ApJ, 766, 95
Line, M. R., Knutson, H., Wolf, A. S., & Yung, Y. L. 2014, ApJ, 783, 70
Line, M. R., Liang, M. C., & Yung, Y. L. 2010, ApJ, 717, 496
Line, M. R., & Parmentier, V. 2016, ApJ, 820, 78
Line, M. R., Vasisht, G., Chen, P., Angerhausen, D., & Yung, Y. L. 2011,
ApJ, 738, 32
Line, M. R., & Yung, Y. L. 2013, ApJ, 779, 3
Line, M. R., Zhang, X., Vasisht, G., et al. 2012, ApJ, 749, 93
Line, M. R., Wolf, A. S., Zhang, X., et al. 2013, ApJ, 775, 137
Luszcz-Cook, S. H., & de Pater, I. 2013, Icarus, 222, 379
Madhusudhan, N., & Seager, S. 2011, ApJ, 729, 41
Mandel, K., & Agol, E. 2002, ApJ, 580, L171
Maness, H. L., Marcy, G. W., Ford, E. B., et al. 2007, PASP, 119, 90
Marley, M. S., Gelino, C., Stephens, D., Lunine, J. I., & Freedman, R. 1999,
ApJ, 513, 879
Marley, M. S., & McKay, C. P. 1999, Icarus, 138, 268
Marley, M. S., Saumon, D., Guillot, T., et al. 1996, Science, 272, 1919
Marley, M. S., Seager, S., Saumon, D., et al. 2002, ApJ, 568, 335
McKay, C. P., Pollack, J. B., & Courtin, R. 1989, Icarus, 80, 23
Morello, G., Waldmann, I. P., Tinetti, G., et al. 2015, ApJ, 802, 117
Morley, C. V., Fortney, J. J., Kempton, E. M.-R., et al. 2013, ApJ, 775, 33
Morley, C. V., Fortney, J. J., Marley, M. S., et al. 2012, ApJ, 756, 172
-. 2015, ApJ, 815, 110
Moses, J. I., Fouchet, T., Bézard, B., et al. 2005, Journal of Geophysical
Research (Planets), 110, 8001
Moses, J. I., Line, M. R., Visscher, C., et al. 2013, ArXiv e-prints,
arXiv:1306.5178
A26
Moses, J. I., Visscher, C., Fortney, J. J., et al. 2011, ApJ, 737, 15
Nettelmann, N., Kramm, U., Redmer, R., & Neuhäuser, R. 2010, A&A, 523,
Pál, A., Bakos, G. Á., Torres, G., et al. 2010, MNRAS, 401, 2665
Parmentier, V., & Guillot, T. 2014, A&A, 562, A133
Pont, F., Gilliland, R. L., Knutson, H., Holman, M., & Charbonneau, D.
2009, MNRAS, 393, L6
Ribas, I., Font-Ribera, A., & Beaulieu, J.-P. 2008, ApJ, 677, L59
Saumon, D., Chabrier, G., & Horn, H. M. V. 1995, Astrophys. J. Suppl., 99,
713
727, 65
27, 2502
Saumon, D., & Marley, M. S. 2008, ApJ, 689, 1327
Shabram, M., Fortney, J. J., Greene, T. P., & Freedman, R. S. 2011, ApJ,
Stamnes, K., Tsay, S.-C., Jayaweera, K., & Wiscombe, W. 1988, Appl. Opt.,
Stevenson, K. B., Harrington, J., Nymeyer, S., et al. 2010, Nature, 464, 1161
Stevenson, K. B., Harrington, J., Lust, N. B., et al. 2012, ApJ, 755, 9
Sudarsky, D., Burrows, A., & Pinto, P. 2000, ApJ, 538, 885
Thompson, S. L. 1990, ANEOS-Analytic Equations of State for Shock
Physics Codes, Sandia Natl. Lab. Doc. SAND89-2951
Thorngren, D. P., Fortney, J. J., & Lopez, E. D. 2015, ArXiv e-prints,
arXiv:1511.07854
Toon, O. B., McKay, C. P., Ackerman, T. P., & Santhanam, K. 1989, Journal
of Geophysical Research, 94, 16287
Toon, O. B., Pollack, J. B., & Sagan, C. 1977, Icarus, 30, 663
Visscher, C., Lodders, K., & Fegley, Jr., B. 2010, ApJ, 716, 1060
von Braun, K., Boyajian, T. S., Kane, S. R., et al. 2012, ApJ, 753, 171
Wong, I., Knutson, H. A., Lewis, N. K., et al. 2015, ApJ, 811, 122
Yung, Y. L., Allen, M., & Pinto, J. P. 1984, ApJS, 55, 465
Zahnle, K. J., & Marley, M. S. 2014, ApJ, 797, 41
Zellem, R. T., Lewis, N. K., Knutson, H. A., et al. 2014, ApJ, 790, 53
Zhang, K., & Hamilton, D. P. 2008, Icarus, 193, 267
|
1901.03074 | 1 | 1901 | 2019-01-10T09:39:40 | VIRTIS-H observations of comet 67P's dust coma: spectral properties and color temperature variability with phase and elevation | [
"astro-ph.EP"
] | We analyze 2-5 micrometre spectroscopic observations of the dust coma of comet 67P/Churyumov-Gerasimenko obtained with the VIRTIS-H instrument onboard Rosetta from 3 June to 29 October 2015 at heliocentric distances r_h = 1.24-1.55 AU. The 2-2.5 micrometre color, bolometric albedo, and color temperature are measured using spectral fitting. Data obtained at alpha = 90{\deg} solar phase angle show an increase of the bolometric albedo (0.05 to 0.14) with increasing altitude (0.5 to 8 km), accompanied by a possible marginal decrease of the color and color temperature. Possible explanations include the presence in the inner coma of dark particles on ballistic trajectories, and radial changes in particle composition. In the phase angle range 50-120{\deg}, phase reddening is significant (0.031 %/100 nm/{\deg}), for a mean color of 2 %/100 nm at alpha = 90{\deg}, that can be related to the roughness of the dust particles. Moreover, a decrease of the color temperature with decreasing phase angle is also observed at a rate of ~ 0.3 K/{\deg}, consistent with the presence of large porous particles, with low thermal inertia, and showing a significant day-to-night temperature contrast. Comparing data acquired at fixed phase angle (alpha = 90{\deg}), a 20% increase of the bolometric albedo is observed near perihelion. Heliocentric variations of the dust color are not significant in the analyzed time period. Measured color temperatures are varying from 260 to 320 K, and follow a r^0.6 variation in the r_h = 1.24-1.5 AU range, close to the expected r_h^0.5 value. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. dustprop-VR-accepted
January 11, 2019
c(cid:13)ESO 2019
VIRTIS-H observations of comet 67P's dust coma: spectral
properties and color temperature variability with phase and
elevation.
D. Bockelée-Morvan1, C. Leyrat1, S. Erard1, F. Andrieu1, F. Capaccioni2, G. Filacchione2, P.H. Hasselmann1, J.
Crovisier1, P. Drossart1, G. Arnold3, M. Ciarniello2, D. Kappel3, A. Longobardo2, M.-T. Capria2, M.C. De Sanctis2, G.
Rinaldi2, and F. Taylor4
9
1
0
2
n
a
J
0
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
4
7
0
3
0
.
1
0
9
1
:
v
i
X
r
a
1 LESIA, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Université, Univ. Paris-Diderot, Sorbonne Paris Cité,
5 place Jules Janssen, 92195 Meudon, France
e-mail: [email protected]
2 INAF-IAPS, Istituto di Astrofisica e Planetologia Spaziali, via del fosso del Cavaliere, 100, 00133, Rome, Italy
3 Institute for Planetary Research, Deutsches Zentrum für Luft- und Raumfahrt (DLR), Rutherfordstrasse 2, 12489, Berlin, Germany
4 Departement of Physics, Oxford University, Parks Rd, OX13PJ, Regno Unito, Oxford, UK
Received ; accepted
ABSTRACT
We analyze 2 -- 5 µm spectroscopic observations of the dust coma of comet 67P/Churyumov-Gerasimenko obtained with the VIRTIS-
H instrument onboard Rosetta from 3 June to 29 October 2015 at heliocentric distances rh = 1.24 -- 1.55 AU. The 2 -- 2.5 µm color,
bolometric albedo, and color temperature are measured using spectral fitting. Data obtained at α = 90◦ solar phase angle show an
increase of the bolometric albedo (0.05 to 0.14) with increasing altitude (0.5 to 8 km), accompanied by a possible marginal decrease of
the color and color temperature. Possible explanations include the presence in the inner coma of dark particles on ballistic trajectories,
and radial changes in particle composition. In the phase angle range 50 -- 120◦, phase reddening is significant (0.031 %/100 nm/◦), for
a mean color of 2 %/100 nm at α = 90◦, that can be related to the roughness of the dust particles. Moreover, a decrease of the color
temperature with decreasing phase angle is also observed at a rate of ∼ 0.3 K/◦, consistent with the presence of large porous particles,
with low thermal inertia, and showing a significant day-to-night temperature contrast. Comparing data acquired at fixed phase angle
(α = 90◦), a 20% increase of the bolometric albedo is observed near perihelion. Heliocentric variations of the dust color are not
significant in the analyzed time period. Measured color temperatures are varying from 260 to 320 K, and follow a r0.6
h variation in the
rh = 1.24 -- 1.5 AU range, close to the expected r0.5
Key words. comet: general -- comets: individual: 67P/Churyumov-Gerasimenko -- infrared : planetary systems
h value.
1. Introduction
The Rosetta mission of the European Space Agency accompa-
nied comet 67P/Churyumov-Gerasimenko (hereafter 67P) be-
tween 2014 and 2016 as it approached perihelion (13 August
2015) and receded from the Sun. Several in situ intruments on
the Rosetta orbiter were dedicated to the study of the phys-
ical and chemical properties of the dust particles released in
the coma. The Micro-Imaging Dust Analysis System (MIDAS,
Riedler et al. 2007) acquired the 3D topography of 1 to 50 µm
sized dust particles with resolutions down to a few nanometers,
and showed that dust particles are agglomerates at all scales
with the smallest subunit sizes of less than 100 nm (Bentley et
al. 2016). A highly porous fractal-like aggregate with a fractal
dimension D f = 1.7 was collected (Mannel et al. 2016). The
Cometary Secondary Ion Mass Analyzer (COSIMA, Kissel et
al. 2007) collected dust particles to image them at a resolution
of 14 µm and performed secondary ion mass spectroscopy. Both
porous aggregates and more compact particles were observed
(Langevin et al. 2016; Merouane et al. 2016). The chemical anal-
ysis indicates that these particles are made of 50% organic mat-
ter in mass, mixed with mineral phases that are mostly anhy-
drous (Bardyn et al. 2017). Carbon is mainly present as macro-
molecular material and shows similarities with the Insoluble Or-
ganic Matter (IOM) found in carbonaceous chondrites (Fray et
al. 2016). The Grain Impact Analyzer and Dust Accumulator
(GIADA, Colangeli et al. 2007) measured the scattered light,
speed, and momentum of individual particles in the size range
of typically 150 -- 800 µm. The majority of the detected dust is
described to be porous agglomerates with a mean density of
785+520−115 kg m−3 (Fulle et al. 2017). GIADA also detected very
low density, fluffy agglomerates, with properties similar to the
MIDAS fractal particles (Fulle et al. 2016).
The remote sensing instruments onboard Rosetta provide
complementary information on the dust properties by measuring
scattered light or thermal emission from particles. From multi-
color imaging using the Optical, Spectroscopic, and Infrared
Remote Imaging System (OSIRIS, Keller et al. 2007), spectral
slopes were measured both for individual particles (Frattin et al.
2017) and for the unresolved dust coma (Bertini et al. 2017).
The observed reddening (e.g. typically 11 -- 14 %/100 nm at λ =
0.4 -- 0.7 µm for the diffuse coma, Bertini et al. 2017) is char-
acteristic of particles made of absorbing material, e.g. organ-
ics (Kolokolova et al. 2004, and references therein). Individual
grains (sizes in the range of centimeters to decimeters) display
differing spectra, which may be related to variations of the or-
Article number, page 1 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
ganic/silicate ratio and the presence of ice (Frattin et al. 2017).
The spectral slopes measured on individual grains display varia-
tions with heliocentric and nucleocentric distances that could be
related to physical processes in the coma affecting the released
material (Frattin et al. 2017). However, spectrophotometric data
of the diffuse coma obtained with OSIRIS do not show any trend
with heliocentric distance and nucleocentric distance (Bertini et
al. 2017).
In this paper, we analyze 2 -- 5 µm spectra of continuum radi-
ation from the dust coma acquired with the high spectral reso-
lution channel of the Visible InfraRed Thermal Imaging Spec-
trometer (VIRTIS-H) onboard Rosetta (Coradini et al. 2007).
This paper is a follow-up of previous work published by Ri-
naldi et al. (2016, 2017) (VIRTIS-M data) and Bockelée-Morvan
et al. (2017) (VIRTIS-H data). Whereas those studies provided
information on the scattering and thermal properties of 67P's
quiescent dust coma at a few dates, namely March-April 2015
and September 2015, we analyze here a comprehensive set of
VIRTIS-H data acquired from June to October 2015 (encom-
passing perihelion on 13 August 2015), with the heliocentric
distance spanning rh = 1.24 -- 1.55 AU. We derive the bolomet-
ric albedo and color temperature of the dust coma, as well as the
spectral slope between 2 -- 2.5 µm following Gehrz & Ney (1992).
These parameters, which have been measured for several comets,
depend on the size distribution, porosity, and composition of the
dust particles (Kolokolova et al. 2004). Measurements obtained
one month after perihelion at rh = 1.3 AU and 90◦ phase angle
are consistent with values measured for most comets (Bockelée-
Morvan et al. 2017). In this paper, we seek for possible variations
with heliocentric distance, altitude, and phase angle.
Section 2 presents the data set. The spectral analysis is de-
scribed in Sect. 3. Results are given in Sect. 4. A discussion
on the observed trends with phase angle and altitude follows
in Sect. 5. Appendix B presents expected thermal properties of
dust particles, and the model used to interpret the variation of the
color temperature with phase angle.
2. The VIRTIS-H data set
VIRTIS is composed of two channels: VIRTIS-M, a spectro-
imager with a visible (VIS) (0.25 -- 1 µm) and an infrared (IR) (1 --
5 µm) channel operating at moderate spectral resolution (λ/∆λ =
70-380), and VIRTIS-H, a cross-dispersing spectrometer provid-
ing spectra with higher spectral resolution capabilities (λ/∆λ =
1300-3500) in eight orders of diffraction covering the range 1.9 --
5.0 µm (Table 1) (Drossart et al. 2000; Coradini et al. 2007). The
infrared channel of VIRTIS-M underwent a cryocooler failure at
the beginning of May 2015. After this date, infrared data were
only collected with VIRTIS-H, and we are focusing on these
data.
As for most Rosetta instruments, the line of sight of VIRTIS-
H is along the Z-axis of the spacecraft (S/C). The instantaneous
field of view (FOV) of this point instrument is 0.58 × 1.74 mrad2
(the larger dimension being along the Y axis). Details on the cal-
ibration process are given in Bockelée-Morvan et al. (2016). The
version of the calibration pipeline is CALIBROS -- 1.2 -- 150126.
VIRTIS-H acquired data cubes of typically 3 h duration in
various pointing modes. For coma studies, the main observing
modes were: 1) limb sounding at a given distance from the comet
surface along the comet-Sun line; 2) limb sounding at a few
stared positions along the comet-Sun line; 3) limb sounding at a
few altitudes and azimuthal angles with respect to the comet-Sun
direction; 4) raster maps (see examples in Bockelée-Morvan et
Article number, page 2 of 13
al. 2016). The data used in this paper were obtained with point-
ing modes 1 -- 3. Dust continuum maps obtained from rasters will
be the topic of a future paper.
We considered data cubes acquired from MTP016/STP058
to MTP024/STP089 covering dates from 30 May 2015 (rh = 1.53
AU) to 30 December 2015 (rh = 2.01 AU), that is, from 74 days
before perihelion to 139 days after perihelion. In total 141 data
cubes were used, though those acquired after 29 October 2015
turned out to be not appropriate for model fitting of the dust con-
tinuum due to low signals (see below). Spectra were obtained by
co-adding acquisitions in the coma for which the exposure time
was typically 3 s. Since we were interested in studying whether
spectral characteristics vary with nucleus distance, we co-added
acquisitions by ranges of tangent altitude (hereafter referred as
to the elevation) with respect to the nucleus surface. This was
done when the signal-to-noise ratio was high enough, and when
the elevation significantly varied during the acquisition of the
data cube (i.e., for pointing modes 2 -- 3). In total 222 spectra
were studied. Figure 1 provides information for each of these
spectra regarding the heliocentric distance, the S/C distance to
nucleus center (∆), and the S/C-nucleus-Sun angle (referred to
as the phase angle). The mean elevation for these spectra is be-
tween 0.8 to 21 km, with 64% of the spectra in the 0.8 -- 4 km
range, and 30% of the spectra in the 4 -- 10 km range (Fig. 1).
For 83% of the spectra, the co-added acquisitions where taken at
elevations which differ by less than 0.5 km. For the rest of the
spectra, elevations of individual acquisitions differ by less than
1.5 km. For stared limb pointing, variation of elevation with time
is observed due to the mutual effects of the complex shape of the
67P rotating nucleus and S/C motion. To define the elevation at
which the spectra refer to, we took the weighted mean of the el-
evation value of each acquisition, with the weight equal to 1/ρ,
where ρ is the distance to nucleus center and is taken equal, for
simplicity, to the elevation plus the mean radius of 67P nucleus
of 2 km. Indeed column densities are expected to vary with a law
close to 1/ρ, so we expect a larger contribution to the signal from
acquisitions with a line-of-sight closer to the nucleus.
Since the VIRTIS-H faint coma signals are affected by stray
light coming from the nearby nucleus, a specific strategy has
been implemented to manage these effects. Stray light polluted
the low wavelength range of each order, and more significantly
order numbered 0, covering the 4 -- 5 µm range (Table 1). Data
cubes obtained at small elevations are the most affected by stray
light. An algorithm developed for stray light removal was ap-
plied (Andrieu et al., in preparation). However, in some cases,
the algorithm was not able to remove all the stray light, espe-
cially in order 0. Therefore, the different orders (which overlap
in wavelength coverage) were merged by selecting the sections
of the orders which are not significantly affected by stray light.
The selected wavelength ranges for each order (Table 1) allow
to reassemble the entire spectrum in the 2 -- 5 µm range. How-
ever, the 4.2 -- 4.5 µm section of order 0 is affected by stray light
but also by the presence of CO2 fluorescence emissions, so it
will not be considered for the analysis of the dust continuum
radiation. The degree of stray light pollution was estimated by
computing the excess of radiance in order 0 with respect to or-
der 1 at wavelengths where these orders overlap (∼ 4.2 µm). We
excluded spectra where this excess is larger than 40%.
At the junction of the selected ranges of the different orders,
spectra with low signal-to-noise ratio show intensity discrepan-
cies to varying degrees, due to the instrumental response which is
low on the edges of each order. These defects were found to lead
to inaccurate results when performing model fitting of the dust
continuum radiation. We defined a criterion based on the ratio
Bockelée-Morvan et al.: 67P dust coma
Table 1. VIRTIS-H diffraction orders and selected wavelengths.
order
λ coverage
selected λ range
(µm)
4.049 -- 5.032
3.477 -- 4.325
3.044 -- 3.774
2.703 -- 3.368
2.432 -- 3.077
2.211 -- 2.755
2.024 -- 2.526
1.871 -- 2.331
4.199 -- 5.000a
3.751 -- 4.216
3.340 -- 3.751
3.047 -- 3.344
2.736 -- 3.047
2.506 -- 2.736
2.316 -- 2.507b
2.005 -- 2.311c
0
1
2
3
4
5
6
7
a 4.2 -- 4.5 µm range excluded for spectral fitting.
b 2.3 -- 2.38 µm range excluded for spectral fitting.
c 2.08 -- 2.17 µm range excluded for spectral fitting.
of the flux measured at 3 µm in order 4 with respect to the value
measured at the same wavelength in order 3. In the initial sample
of 222 spectra, this ratio, referred as to T ES T3.0, varies between
0.95 to 2.7, with a value close to 1 indicating a high quality spec-
trum. We only considered model fitting results for spectra com-
plying T ES T3.0 < 1.35 (173 spectra among the 222). In the fol-
lowing sections, we will also discuss results obtained for the best
quality spectra fulfilling T ES T3.0 < 1.1 (49 spectra). For these
high-quality spectra, the signal-to-noise ratio (SNR) at 4.65 µm
is in the range 30 -- 80, with a few exceptions (the relevant root
mean square is computed on the spectrum from the statistics of
the residuals between the observed spectrum and model fit (Sect.
3) in the range 4.5 -- 4.8 µm). The SNRs at 3.3 µm (order 3) and
2.3 µm (order 7) are a factor 3 -- 4 lower. Spectra with SNR at
4.65 µm less than 12 were not considered. After excluding spec-
tra with high stray light pollution, at the end, 99 (respectively
49) spectra complying T ES T3.0 < 1.35 (respectively T ES T3.0
< 1.1) where found appropriate for model fitting. The covered
time period is -- 71 to +78 d with respect to perihelion (3 June
to 29 Oct. 2015, rh = 1.24 to 1.55 AU). The best quality spectra
cover dates from -- 43 to +78 d. Table A.1 provides information
on these 99 spectra, such as VIRTIS-H observation identification
number, start time of the data cube, date with respect to perihe-
lion, spacecraft distance to nucleus center, heliocentric distance
and phase angle.
Figure 2 shows two examples of high quality spectra affected
by negligible stray light, obtained for the coma on 22 Jul. 2015
and 8 Aug. 2015 with SNR of respectively 76 and 50 at 4.65 µm.
The dust continuum consists of scattered sunlight at λ < 3.5 µm
and thermal radiation at longer wavelengths. Fluorescence emis-
sion bands of H2O, CO2 and 13CO2 are observed in the 2.5 -- 3.0
µm and 4.2 -- 4.4 µm ranges (see detailed description in Bockelée-
Morvan et al. 2015, 2016).
3. Model fitting
In order to analyze the dust continuum radiation, we followed
the approach presented by Bockelée-Morvan et al. (2017) which
consists in modeling the dust spectrum as the sum of scattered
solar flux and thermal emission (described by a blackbody func-
tion). The free parameters of the model fitting are the color tem-
perature Tcol, the spectral index of the reflectance, which allows
us to derive the dust color S (cid:48)
col in the 2.0 -- 2.5 µm range, and the
bolometric albedo A(θ), where θ is the scattering angle (hereafter
we will use instead the phase angle α = 180◦ − θ, and assimilate
the phase angle to the S/C -- Comet -- Sun angle, which is a good
Fig. 1. Geometrical informations for the constituted data set of averaged
limb spectra. From top to bottom: S/C-nucleus-Sun angle (referred as to
the phase angle), S/C distance to comet center (∆), heliocentric distance
(rh), and mean elevation (H) of the line-of-sight.
Fig. 2. VIRTIS-H spectra of comet 67P obtained at different elevations
above the surface. Cube T1_00396220410 acquired on 22 Jul. 2015
(rh = 1.27 AU) with a mean elevation H = 1.4 km (red line). Cube
T1_00397139303 acquired on 8 Aug. 2015 (rh = 1.25 AU) using se-
lected acquisitions with H = 6.2 km (black line). The phase angle is
∼89◦ for both spectra. The model fits to the continuum are shown in
cyan, with derived parameters (Tcol(K), A, S (cid:48)
col(% per 100 nm)) = (295,
0.07, 2.3) and (289, 0.10, 2.2) for the 22 Jul. and 8 Aug. spectra, respec-
tively. The spectra fulfill the quality criterion T ES T3.0 < 1.1.
approximation given the large S/C distance to the comet). From
the color temperature, we can derive the so-called superheating
factor S heat, defined as the ratio of the observed color temper-
ature Tcol to the equilibrium temperature Tequ of a fast rotating
body:
Tcol
Tequ
,
S heat =
with
Tequ = 278r−0.5
h
[K],
(1)
(2)
Article number, page 3 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
only consider statistical noise and not defects in the spectra re-
lated, e.g., to the calibration, or possible residual stray light (see
Sect. 2). For example, for the fit shown in Fig. 3, 1-σ uncertain-
ties are 0.3%, 1%, and 4% for Tcol(K), A and S (cid:48)
col, respectively
(1-σ confidence levels were derived as explained in Bockelée-
Morvan et al. 2017). However, though the noise level is low be-
tween 4.5 and 5 µm (SNR = 76), the fit is not fully satisfactory in
this spectral region (Fig. 3). There is also a small radiance offset
at 3.752 µm, which corresponds to the junction of the selected
wavelength ranges in orders 1 and 2 (Table 1).
It is important to point out that the retrieved free parameters
are somewhat correlated. This is because scattered light and ther-
mal emission contribute both to the continuum in a significant
fraction of the 2 -- 5 µm spectrum (Fig. 3). A statistical analysis
based on contours of equal χ2 shows that Tcol and S (cid:48)
col (and con-
sequently A) are correlated among them. Dust color and color
temperature are negatively correlated, whereas the bolometric
albedo and color temperature are positively correlated. As a re-
sult, significant flaws somewhere in the spectrum can lead to
spurious results which follow this trend (e.g., a lower Tcol com-
bined with higher S (cid:48)
col, and lower A). Effectively, we observed
that spectra fulfilling the quality test T ES T3.0 > 1.1 have lower
Tcol, combined with higher S (cid:48)
col and lower A, compared to val-
ues retrieved for higher quality spectra with T ES T3.0 < 1.1. This
will be further discussed in Sect. 4.1.
4. Results
Figure 4 shows the bolometric albedo, color, color temperature
and superheating factor as a function of date with respect to per-
ihelion, for the 99 spectra with T ES T3.0 < 1.35 and stray light
excess < 1.4, as explained in Sect. 2. The different points also
characterize the 2 -- 5 µm dust emission at various elevations of
the line of sight (as indicated by the color code), and phase an-
gles (cf overplotted phase information on Fig. 4A, C, D). We
recall that the elevation H corresponds to the altitude of the tan-
gent point (Sect. 2). The results are also listed in Table A.1.
Tcol ranges from 260 K to 320 K, and follows approximately
the r−0.5
variation expected from the balance between absorbed
h
solar radiation and radiated thermal energy (Fig. 4B). The su-
perheating factor S heat is typically 1.2 before perihelion (phase
angle α of about 90◦). However, strong variations of S heat are
observed after perihelion when the Rosetta S/C was flying out of
terminator (with α on the order of 60◦ or reaching 120◦). These
variations seem to be correlated with changes in the phase angle
(Fig. 4C). A strong correlation with phase angle is also observed
for the color S (cid:48)
col (Fig. 4D). Whereas S heat decreases with in-
creasing phase angle, the reverse is observed for the color. As
for the bolometric albedo, higher values are measured after peri-
helion (Fig. 4A), which is consistent with the phase function of
cometary dust, which has a "U" shape with a minimum at α =
90 -- 100◦ (Bertini et al. 2017). However, a trend for higher albe-
dos at higher elevations and/or near perihelion is also suggested
(Fig. 4A).
In the subsequent subsections, we will analyze eleva-
tion/time and phase variations of Tcol, A(θ) and S (cid:48)
col. We will
also study the intensity ratio between scattered light and thermal
emission. The reference for scattered light is the radiance mea-
sured at λ = 2.44 µm, obtained from the median of the radiances
between 2.38 and 2.5 µm (order 6, Table 1). For the thermal
emission, the reference is the radiance at λ = 4.6 µm (median of
radiances between 4.5 -- 4.7 µm). The intensity ratio fscatt/ ftherm
is obtained by ratioing radiances in units of W m−2 sr−1 Hz−1.
At constant phase angle, if the dust size distribution and com-
Fig. 3. Example of model fit. The spectrum in black is a VIRTIS-H
spectrum of comet 67P obtained on 22 Jul. 2015 (T1_00396220410),
with the regions excluded for fitting or presenting water and CO2 emis-
sion lines not shown (the full spectrum is given in Fig. 2). The model fit
to the continuum, which corresponds to the sum of scattered light (plain
orange line) and thermal radiation (dashed green line) is shown in red.
Retrieved parameters are Tcol = 295 ± 1 K (corresponding to S heat =
1.194 ± 0.003), A = 0.068 ± 0.001, S (cid:48)
= 2.3 ± 0.1 % per 100 nm.
col
where rh is in AU (this unit is used throughout the paper).
The definitions of A(θ) and S heat follow the prescription of
Gehrz & Ney (1992), which allowed us to compare 67P's dust
infrared emission properties to other comets for which these pa-
rameters have been measured (Bockelée-Morvan et al. 2017).
The bolometric albedo A(θ) is approximately equal to the ratio
between the scattered energy by the coma to the total incident
energy, and scales proportionally to the geometric albedo times
the phase function. Further details can be found in Bockelée-
Morvan et al. (2017).
The dust color (or reddening) is measured in %/100 nm using
the dust reflectance at 2.0 µm and 2.5 µm:
S (cid:48)
col
scatt(2.5µm) − Rfit
= (2/500) × Rfit
Rfit
scatt(2.5µm) + Rfit
scatt(2.0µm)
scatt(2.0µm)
,
(3)
where Rfit
length λ divided by the solar flux at λ (Kurucz et al. 1992).
scatt(λ) is the fitted scattered light (Fig. 3) at the wave-
In the fitting process, the spectral region 4.2 -- 4.5 µm show-
ing CO2 emissions and stray light was masked. However, unlike
in Bockelée-Morvan et al. (2017), the 3.3 -- 3.6 µm region was
kept, as only very faint emission features from organics are ob-
served in this region (Bockelée-Morvan et al. 2016). The model
includes a synthetic H2O fluorescence spectrum (described in
Bockelée-Morvan et al. (2015) with a rotational temperature of
100 K) with the total intensity used as free parameter, so the
2.5 -- 3.0 µm region presenting water lines could be considered.
Despiking (using median filtering) was applied, removing spikes
such as those seen in the spectrum of Fig. 3, though this was
not found critical. We checked that the fitting method, which
uses the χ2 minimization algorithm of Levenberg-Marquardt,
provides correct results by applying it to synthetic spectra to
which synthetic noise resembling the noise present in 67P spec-
tra was added. Applying our algorithm to the data set presented
in Sect. 2, the best fits have a reduced χ2 very close to 1 (0.94 on
average).
Figure 3 shows an example of a model fit to a high-SNR dust
spectrum, with the two components, scattered light and thermal
emission, shown separately, and the retrieved free parameters
indicated in the caption. The uncertainties in the retrieved pa-
rameters are probably somewhat underestimated because they
Article number, page 4 of 13
Bockelée-Morvan et al.: 67P dust coma
Fig. 4. Bolometric albedo (A), color temperature (B), superheating factor (C) and dust color (D) as a function of time with respect to perihelion.
The color of the symbols is according to elevation (as indicated in the colorbar). The phase angle is plotted as a dashed line in plots A, C, and D.
The dashed line in plot B) is the color temperature obtained for a value of the superheating factor of 1.19 and a heliocentric variation following
r−0.5
h
(Eq. 1). The fitted spectra fulfill the condition TEST3.0 < 1.35.
h
position do not vary with time and in the coma, this ratio is ex-
pected to only exhibit a heliocentric dependence proportional to
r−2
h /BB(Tcol), where BB is the blackbody function at Tcol which
varies as r−0.5
. We corrected the derived intensity ratios from
this heliocentric dependence assuming S heat = 1.2 and converted
it to the value at 1 AU ( fscatt/ ftherm(1AU)). As discussed at the
end of Sect. 3, spectral fitting to spectra presenting some offsets
at the junction of the orders can provide inaccurate results. On
the other hand, the intensity ratio fscatt/ ftherm(1AU) is directly
measured on the spectra, and provides reliable trends.
4.1. Results at 90◦ phase angle
In this Section, we only consider measurements obtained at
phase angles between 83◦ and 90 ◦ (mean value of 89◦). These
data were acquired mainly before perihelion. The color tempera-
ture follows Tcol = (338 ± 1)r−0.60±0.01
K in the heliocentric range
rh 1.24 -- 1.5 AU. Considering only the best quality data (covering
1.24 -- 1.34 AU), one finds Tcol = (333 ± 3)r−0.51±0.03
K.
h
h
Figure 5 shows the bolometric albedo, color, and superheat-
ing factor as a function of elevation H (and rh using a color
gradient for the symbols). The results from the highest quality
spectra (T ES T3.0 < 1.1) are shown with squares, whereas the
other data (1.1 < T ES T3.0 < 1.35) are shown with dots. S heat
col have mean values of 1.19±0.01 and 2.0±0.2 % per 100
and S (cid:48)
nm, respectively. Lower quality spectra show lower S heat and
higher color S (cid:48)
col and albedo values that may be inaccurate (see
Sect. 3). To test this hypothesis, we performed spectral fitting,
fixing the color temperature. We found that an underevaluation
of S heat by 4% (S heat=1.15 instead of 1.2) would decrease the de-
rived albedo by ∼60%. Effectively the albedo derived for the low
quality spectra giving S heat=1.15 is lower by this order of mag-
nitude (upper panel of Fig. 5). So, results from these spectra, es-
pecially those for which the derived color S (cid:48)
col is well above the
mean value, are a priori doubtful. On the other hand, the intensity
ratio fscatt/ ftherm(1AU), which is proportional to the bolometric
albedo, presents a similar behavior with elevation and heliocen-
tric distance though discrepancies between high quality and low
quality spectra are somewhat smaller (Fig. 6). In conclusion, the
trend for an enhanced albedo at low heliocentric distance seen in
Fig. 5 is likely real, as well as the trend for increased superheat-
ing with decreasing rh.
A marginal decrease of S heat and S (cid:48)
col with increasing eleva-
tion H is suggested (best data), with a Pearson correlation coeffi-
cient R of -- 0.34 and -- 0.40, respectively (Fig. 5). We performed a
multi regression analysis for studying variations with both rh and
altitude. A weak rh variation in r−0.15±0.05
is suggested for S heat,
which improves the correlation coefficient with altitude to R =
-- 0.55, with S heat ∝ H−0.009±0.003. Multi regression analysis did
not provide convincing results for S (cid:48)
col: no reliable variation of
the color with rh could be identified in this data set. Altogether,
h
Article number, page 5 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
Fig. 6. Variation of fscatt/ ftherm(1AU) (deduced from the ratio of the radi-
ances at 2.44 µm and 4.6 µm, see text) with elevation H for α = 83 -- 90◦.
The color coding and symbols are as for Fig. 5. Data with TEST3.0 <
1.35 are considered. The dashed-dotted and dashed lines correspond to
a power law fit for data obtained between -- 71 d and -- 10 d wrt perihelion
(∝ H+0.31) and between -- 2 d to 21 d (∝ H+0.22), respectively.
Fig. 7. VIRTIS-H spectra of comet 67P obtained at different phase
angles (shown in logarithmic scale). Top spectrum (red): cube
T1_00399392471 acquired on 28 Aug. 2015 (rh = 1.26 AU, α= ∼72◦)
with a mean elevation of 4.1 km. Bottom spectrum (black): cube
T1_00400433767 acquired on 7 Sept. 2015 (rh = 1.29 AU, α= ∼120◦)
with a mean elevation of 2.2 km. The model fits to the continuum are
shown in cyan, with derived parameters (Tcol(K), A, S (cid:48)
col(% per 100 nm),
S heat) = (304, 0.14, 1.3, 1.23) and (283, 0.09, 2.8, 1.15) for 28 Aug. and
7 Sept., respectively. The spectra fulfill the quality criterion T ES T3.0 <
1.1 and do not present significant stray light.
forming a multi regression analysis to fscatt/ ftherm(1AU). Com-
paring data acquired between -- 2 d to 21 d wrt perihelion to those
acquired before (up to end July 1015), an average increase of
20% of fscatt/ ftherm(1AU) (and hence of the albedo) is suggested
(Fig. 6). The variation with elevation follows fscatt/ ftherm(1AU)
∝ H+0.27±0.05, where the power law index is the average of the
indexes obtained for the two time periods (Fig. 6). The bolomet-
ric albedo measured on the high quality spectra follows the same
variation.
Fig. 5. Variation of bolometric albedo, superheating factor and color
with elevation H. Data obtained with phase angle α = 83 -- 90◦ are con-
sidered. The color is a function of the heliocentric distance, as given
by the colorbar. Only data with TEST3.0 < 1.35 are plotted. Those with
T ES T3.0 < 1.1 are shown with large squares. The dashed lines corre-
spond to a power law (for albedo (∝ H−0.39±0.01)) or a linear fit (for su-
perheating factor and color) to the data points with T ES T3.0 < 1.1.
however, variations of S heat and S (cid:48)
AU) are small.
col with H and rh (1.24 to 1.35
There is evidence for a significant increase of the bolomet-
ric albedo with H (Fig. 5). This is illustrated in Fig. 2, which
displays two spectra obtained at H= 1.4 and 6.2 km, the former
showing a lower flux ratio fscatt/ ftherm. Since S heat (or Tcol) shows
weak variation with H, the increase of A with H reflects the in-
crease of fscatt/ ftherm(1AU) with H, shown in Fig. 6. We looked
for possible variations of A with rh or seasonal changes, per-
Article number, page 6 of 13
Bockelée-Morvan et al.: 67P dust coma
4.2. Phase variations
The dust color and color temperature exhibit a strong correlation
with phase angle. The dust color is larger at large phase angles
(Figs 4C). On the other hand, the reverse is observed for the color
temperature, as best seen when looking to the trend followed by
the superheating factor (Fig. 4D). Figure 7 compares two spectra
acquired with a one week interval at α= 72 and 120◦. The ratios
of the thermal emissions in orders 1 (3.7 -- 4.2 µm) and 0 (4.5 -- 5
µm) present subtle differences (by up to 9%) explained by a color
temperature higher by 20 K at low phase. The fitting algorithm
retrieves also a bluer color at low phase to match the 3.0 -- 3.5 µm
radiances.
We present in Figs. 8B and C the variations of the color and
superheating factor with phase angle. To avoid clutter at α = 90◦,
only dates after -- 2 d wrt perihelion are plotted. The phase depen-
dences found using the best quality data are ∼ 0.3 K/◦ for Tcol,
and 0.031 %/100 nm/◦ for the dust color. Significant variations
with elevation are not seen.
The bolometric albedo (measured at 2 µm) follows a phase
variation which matches the phase function measured at 537 nm
by Bertini et al. (2017) during MTP020/STP071 (end August
2015) (Fig. 8A). The VIRTIS data present a large scatter, which
prevents further comparison. Note that the dust phase function is
expected to be wavelength-dependent. The variation of bolomet-
ric albedo with elevation at low phases (α < 80◦) follow a H0.25
law for the best data, similar to the one measured at α = 90◦, but
the data show significant scatter with respect to this variation.
Phase variations of color and color temperature of cometary
dust have never been reported in the literature. From detailed
analysis and multiple checks, we can rule out biases related to
the fitting algorithm and data quality. Since retrieved parame-
ters are somewhat correlated (Sect. 3), another test was to fix the
color temperature according to Eq. 1, with S heat fixed to the α
= 90◦ value of 1.19 (Sect. 4.1). Despite the increase of the χ2
values, the color trend with phase remains. However, the bolo-
metric albedo shows a monotonic slight decrease with decreas-
ing phase angle (i.e., no backscattering enhancement), which is
unexpected from scattering models, thus reassuring us that the
observed phase variations are real.
5. Discussion
In summary, analysis of the dust 2 -- 5 µm continuum radiation
from 67P's coma shows: i) a mean dust color of 2%/100 nm
and superheating factor of 1.19 at 90◦ phase angle, consistent
with previous VIRTIS-H measurements (Bockelée-Morvan et al.
2017); ii) a factor 2.5 increase of the bolometric albedo with
increasing elevation from H = 0.5 to 8 km; iii) an increase of
dust color temperature with decreasing phase angle at the rate of
∼ 0.3 K/◦ ; iv) spectral phase reddening at a rate of 0.032 %/100
nm/◦. More marginally, decreasing color temperature and color
with increasing H are possibly observed, as well as 20% higher
albedo values after perihelion.
Fig. 8. Variation of bolometric albedo A (top, A), superheating factor
S heat (middle, B) and color S (cid:48)
col (bottom, C) with phase angle using data
taken after 11 August 2015. Data are from spectra with TEST3.0 < 1.35
with the best data (T ES T3.0 < 1.1) shown with squares. A) The dashed
line is the phase function measured during MTP020 by Bertini et al.
(2017) (OSIRIS Green filter -- 537 nm). B) The dashed line is a linear
fit to S heat data with a slope of -- 0.0014/◦ ( -- 0.33 K/◦ at rh = 1.35 AU).
The dashed-dotted line is from the Near Earth Asteroid Thermal Model
(NEATM) with beaming parameter η = 1.6, emissivity = 0.9 and rh =
1.35 AU. The solid line is from our dust thermal model (Sec. 5.2) with
thermal parameter Θ = 0.1, optical depth fraction of isothermal particles
fiso = 0.8, fheat = 1.05, and = 0.9. C) The dashed line is a linear fit to
the data giving a phase reddening of 0.032 %/100 nm/◦.
Article number, page 7 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
5.1. Phase reddening
The photometric properties of the dust coma present similarities
with the nucleus surface. 67P's nucleus shows a phase redden-
ing which has been observed both in the optical (VIS) (0.5 -- 0.8
µm) and in the near-IR (1 -- 2 µm) ranges (Ciarniello et al. 2015;
Longobardo et al. 2017; Feller et al. 2016; Fornasier et al. 2016).
In the near IR, 67P's nucleus color is 3.9 %/100 nm at α = 90◦,
with a phase reddening between 0.013 -- 0.018 %/100 nm/◦ (Cia-
rniello et al. 2015; Longobardo et al. 2017). Phase reddening is
higher in the VIS (0.04 to 0.1 %/100 nm/◦), with lower values
near perihelion associated to a bluing of the surface (Fornasier
et al. 2016). For the dust coma, the weighted mean of the VIS
values measured by Bertini et al. (2017) using OSIRIS data (ex-
cluding spurious MTP026 results) yields 0.025%/100 nm/◦. This
is close to the values that we are measuring in the near-IR. How-
ever, it should be kept in mind that the VIS values are from data
with LOS perpendicular to the nucleus -- S/C vector (Bertini et al.
2017), so they pertain to the dust coma in the near-spacecraft
environment, whereas the near-IR values characterise the near-
nucleus coma. There are several lines of evidence that the dust
properties vary with elevation, as discussed later on.
Phase reddening is observed for many Solar System bodies,
including zodiacal light (Leinert et al. 1981). For planetary sur-
faces, phase reddening can be interpreted as an effect of multi-
ple scattering. For dark and porous bodies as 67P, multiple scat-
tering is relevant despite the low albedo thanks to the increase
of scattering surfaces caused by the roughness of the particles
present on the nucleus surface (Schröder et al. 2014). Labora-
tory experiments combined with numerical simulations have in-
deed highlighted the role of microscopic roughness in producing
such a spectral effect (Beck et al. 2012; Schröder et al. 2014).
Particle irregularities at a spatial scale less than the wavelength
are also invoked to explain the phase reddening seen in the vi-
sual for interplanetary dust (10 -- 100 µm sized) (Schiffer 1985).
Then, the phase reddening observed in the 67P coma could be
related to the porous structure of the particles, providing those
contributing to scattered light are sufficiently large. The relative
similarity in the phase curves of the dust coma and surface (es-
pecially the backscattering enhancement) is consistent with the
predominance of large and fluffy dust particles in the coma, as,
e.g., discussed by Moreno et al. (2018), Bertini et al. (2019) and
Markkanen et al. (2018). Other evidence for relatively large (≥
10 µm) scatterers in the coma of 67P include dust tail modeling
(Moreno et al. 2017) and the unexpected low amount of submi-
cron and micron-sized particles collected by the Rosetta's MI-
DAS experiment (Mannel et al. 2017).
5.2. Phase variation of the color temperature
The color temperature excess with respect to the equilibrium
temperature expected for isothermal grains is a common prop-
erty of cometary atmospheres. The superheating factor measured
for 67P of ∼ 1.2 is in the mean of values observed in other comets
(Bockelée-Morvan et al. 2017). This temperature excess is usu-
ally attributed to the presence of submicrometric grains com-
posed of absorbing material (Hanner 2003; Kolokolova et al.
2004). Bockelée-Morvan et al. (2017) showed that this tempera-
ture excess could result from the contribution of hot fractal-like
aggregates to near-IR thermal emission, these particles having in
turn little input to scattered light. In this case, based on Mie mod-
eling, the minimum size of the more numerous and more com-
pact particles would be ≥ 20 µm (Bockelée-Morvan et al. 2017).
The observed decrease of the color temperature with increasing
Article number, page 8 of 13
phase angle can not be explained by variations of the dust size
distribution with solar azimuth angle (Shou et al. 2017), which
would induce a phase curve symmetric with respect to α = 90◦.
On the other hand, this trend can be caused by non-isothermal
grains showing day-to-night thermal contrast. This explanation
holds for Saturn's C-ring whose thermal emission shows varia-
tions with solar phase angle (Altobelli et al. 2008; Leyrat et al.
2008).
To test this hypothesis, in a first approach we used the Near
Earth Asteroids Thermal Model (NEATM) (Harris 1998) for
describing the variation of the temperature over the surface of
comet dust particles. NEATM assumes an idealized non-rotating
spherical object with a temperature decreasing from a maximum
at the subsolar point to zero at the terminator (there is no night-
side emission). For low albedo bodies, the surface temperature at
latitude θ(cid:48) -- π/2 and longitude φ(cid:48) (subsolar point at θ(cid:48) = 90◦ and
φ(cid:48) = 0◦) follows:
NEATM(sin(θ
TNEATM = T SS
with
(cid:48)) cos(φ
(cid:48)))0.25,
T SS
NEATM
=
394
r0.5
h (η)0.25
[K],
(4)
(5)
where is the emissivity (taken equal to 0.9), and η is the so-
called beaming parameter, which is used in asteroid studies as a
calibration coefficient to account for the effects of thermal iner-
tia, rotation and surface roughness. T SS
NEATM is the temperature at
the subsolar point. Thermal emission is calculated considering
the surface elements facing the observer and, therefore, depends
on the phase angle (Harris 1998). We computed NEATM 3 -- 5
µm spectra for a range of phase angles and η values. By fitting a
blackbody to these spectra, we derived color temperatures and,
using Eq. 1, the corresponding superheating factors. The phase
variation of these computed superheating factors (dashed-dotted
curve in Fig. 8B) matches the variation measured for 67P dust,
and the mean observed value S heat = 1.19 at α ∼ 90◦ (Sect. 4.1)
is obtained for η = 1.58. We determined η for each of the data
points shown in Fig. 8B. Inferred η values do not show signif-
icant phase dependence and average out at 1.59 ± 0.17. This
value is intermediate between the limiting cases η = 1 (high day-
to-night contrast due to low thermal inertia, slowly spinning par-
ticles, or spin axis along Sun direction) and η = 4 (isothermal
particles). This suggests that both isothermal and non-isothermal
grains are contributing to 67P dust thermal emission in the 3 -- 5
µm wavelength range.
To go further into the interpretation of the data, we devel-
oped a simple model (Appendix B), considering a bimodal dis-
tribution of dust particles consisting of a mixture of isothermal
particles and particles presenting day-to-night temperature con-
trast. The diurnal temperature profile of non-isothermal particles
is described by the thermal parameter Θ introduced by Spencer
et al. (1989), which depends on their thermal properties (which
is a function of porosity) and spinning rate. Figure 9 shows ex-
amples of diurnal temperature profiles, which computations are
described in Appendix B. Expected Θ values for 67P dust parti-
cles are also given in Appendix B (Fig. B.1). The relative con-
tribution of the isothermal particles to the total optical depth is
parameterized by the quantity fiso (in the range 0 -- 1), and their
physical temperature Tiso is in excess with respect to the equilib-
rium temperature by a factor fheat/0.25. Figure 10 shows the su-
perheating factor and the slope of the phase variation of the color
Bockelée-Morvan et al.: 67P dust coma
Fig. 9. Modelled particles diurnal temperature profiles at equator for
different values of the thermal parameter Θ. The subsolar point is at
latitude zero and at longitude zero. For large Θ, the temperature reaches
the constant isothermal value Tiso. Calculations are for rh = 1.35 AU
and =0.9.
temperature as a function of Θ for different values of fiso, consid-
ering values of fheat of 1.0 and 1.05. The non-monotonic behav-
ior of the phase dependence for low Θ and fiso values is because
the 3 -- 5 µm wavelength range is more sensitive to high temper-
atures (e.g., for high day-to-night temperature contrast, only the
warm surface areas contribute to the brightness). Best match to
the measurements is obtained for Θ ≤ 2, corresponding to a sig-
nificant day-to-night temperature contrast (> 1.5). Such low val-
ues of Θ imply slowly spinning particles with high porosity, low
thermal inertia or non-spinning particles (Fig. B.1). The relative
contribution of isothermal particles is not heavily constrained.
For Θ = 2.0, one finds solutions with fiso = 0.2 -- 0.4. On the other
hand, for Θ = 0.1, a good fit to the data is obtained for fiso = 0.8
(see Fig. 8). We note that a good fit to the color temperature and
its phase variation is obtained providing the physical temperature
of the isothermal grains is in excess by 8% with respect to the
expected equilibrium temperature (i.e., fheat = 1.05, considering
an assumed emissivity of 0.9). Therefore, the presence of non-
isothermal grains with day-side surface temperature well above
the equilibrium temperature, cannot alone explain the superheat-
ing factor observed for cometary dust. As already discussed, a
possible explanation is a significant contribution of submicron-
sized absorbing grains (Hanner 2003) or, alternatively, of highly
porous fractal-like aggregates with sub-micron size monomers,
as these particles can be warmer than more compact particles
(Bockelée-Morvan et al. 2017).
A realistic size distribution of the dust particles is obviously
not bimodal. It is interesting to estimate the critical radius below
which the particles are isothermal, and to compare it to estimated
diurnal skin depths (Appendix B, Fig. B.1). Assuming a power
law for the size distribution (dN ∝ a−βda, where a is the parti-
cle radius), this critical radius acrit depends on the size index β
and minimum and maximum sizes of the particles, amin and amax,
and can be computed using the inferred relative contribution to
the total optical depth of the two populations of particles ( fiso
and 1 -- fiso) (see equations in Leyrat et al. 2008). acrit increases
with increasing amin, amax and fiso, and with decreasing β. Let us
consider size ranges amin = 1 -- 20 µm and amax = 1 -- 10 cm, con-
Fig. 10. Superheating factor S heat (A) and phase linear dependence of
the color temperature (B) from the thermal model parameterized by the
thermal parameter Θ, the optical depth contribution of isothermal par-
ticles fiso and their physical temperature Tiso. The shaded areas corre-
spond to the measurements in the 67P coma.
sistent with constraints obtained for 67P dust (Bockelée-Morvan
et al. 2017; Mannel et al. 2017; Ott et al. 2017; Schloerb et al.
2017; Moreno et al. 2018; Markkanen et al. 2018). For fiso = 0.8
(solution obtained for Θ = 0.1) and β = 2.5 (respectively β = 3.0),
acrit is in the range 0.6 to 6 cm (respectively 0.15 -- 1.7 cm). These
values of acrit are on the order of or larger than the estimated
diurnal skin depths of ∼ 0.3 cm for slowly spinning and high
porosity particles with Θ = 0.1 (Appendix B, Fig. B.1). For fiso
= 0.3 (solution obtained for Θ = 2.0), acrit is in the range 0.09 to
0.9 cm (respectively 0.0015 -- 0.02 cm) for β = 2.5 (respectively β
= 3.0). For particles with porosity of 0.5 -- 0.9 and spinning rates
consistent with Θ = 2.0, we expect diurnal skin depths from 0.01
to 0.3 cm. Altogether, except for size distributions with the opac-
ity dominated by small particles (those with β = 3 and amin < 10
µm, or β > 3), we infer that the critical particle size separating
isothermal and non-isothermal particles is on the order or larger
than the diurnal skin depth. This is a satisfactory result since we
expect particles with size smaller than the diurnal skin depth to
be isothermal due to internal heat transfer.
Article number, page 9 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
5.3. Radial variation of the bolometric albedo
Measured bolometric albedos for 67P quiescent dust coma range
between 0.05 and 0.15 at 90◦ phase angle (Fig. 5) and encom-
passe values measured in other comets, as previously discussed
by Bockelée-Morvan et al. (2017). These values correspond to a
low geometric albedo, and is consistent with dust particles made
of dark material (Kolokolova et al. 2004; Bockelée-Morvan et
al. 2017, and references therein). The VIRTIS-H observations
suggest an increase of the dust bolometric albedo with increas-
ing radial distance (Sect. 4). Albedo maps obtained for comets
1P/Halley and 21P/Giacobini-Zinner by combining visible light
and thermal infrared images show a similar trend: the albedos in-
crease radially from the nucleus, except along the tail, where the
albedos are smaller (Telesco et al. 1986; Hammel et al. 1987).
Variations in albedo may result from different composition, par-
ticle size, shape and structure. For example, large fluffy grains
may have reduced albedos because they induce multiple scatter-
ing events that allow more light to be absorbed. For this reason,
the lower albedos near the nucleus and in the tail of comets 1P
and 21P have been interpreted as due to the presence of large,
fluffy grains escaping the nucleus with low velocities and con-
fined in the orbital planes of the comets (Telesco et al. 1986;
Hammel et al. 1987). We may thus invoke an enhanced pro-
portion of chunks in the inner coma of 67P, in line with the
conclusion obtained by Bertini et al. (2019) from the variation
of the backscattering enhancement with nucleocentric distance.
During the perihelion period, comet 67P underwent numerous
outbursts (Vincent et al. 2016), which likely populated the in-
ner coma with large, slowly-moving dust particles, as observed
for comet 17P/Holmes after its massive 2007 outburst (Reach et
al. 2010; Boissier et al. 2012). In addition, evidence for parti-
cles falling back to the nucleus are several (Keller et al. 2017).
Models of the density distribution for a coma dominated by grav-
itationally bound particles on ballistic trajectories predict an ex-
cess of particles in the inner coma with respect to the density
expected for free-radial outflow (Chamberlain & Hunten 1987;
Gerig et al. 2018). There are some hints of such a deviation from
free-radial outflow in OSIRIS optical images (Gerig et al. 2018),
which would be amplified if the observed trend for a smaller
albedo at smaller cometocentric distances is considered. Devi-
ations are also conspicuous for the dust thermal radiation mea-
sured in the microwave, which samples essentially large parti-
cles, and shows a steep decrease of the column density at impact
parameters below 10 km (Schloerb et al. 2017).
Whereas we argued previously for radial variations of the
optical properties of the individual grains, changes in the par-
ticle size distribution may also affect the bolometric albedo of
the coma. Anomalously high bolometric albedos were measured
in the very active comet C/1995 O1 (Hale-Bopp), and during
strong jet activity of 1P/Halley, which were found to be corre-
lated with a high silicate 10µm-band contrast, and a high su-
perheating factor S heat, suggesting that the presence of a large
amount of small particles was responsible for these high albe-
dos (Tokunaga et al. 1986; Mason et al. 2001; Hanner 2003).
Similarly, the rapidly moving 67P outburst ejecta displayed high
A and S heat, together with blue colors, characteristics of small
particles (Bockelée-Morvan et al. 2017). Mie calculations for a
porous mixture of olivine and amorphous carbon at 90◦ phase
angle predict an increase of A from a value of 0.05, when only
particle sizes > 1 µm are considered, to values up to 0.20 when
submicron particles are present. However, the increase of A is
expected to be correlated with an increase of the superheating
factor. This trend between A and S heat with increasing elevation
Article number, page 10 of 13
is not observed (Sect. 4, Fig. 5). Therefore, dust fragmentation
is likely not responsible for the increase of A with elevation.
Changes in the albedo may also be related to a change in the
particle composition. Particles made of less absorbing material
are expected to be brighter, cooler and bluer. This trend is ob-
served with increasing elevation, which would then imply that
evaporation of some dark material took place in the inner coma.
Evidence for the degradation of grains in the coma of 67P are
however still very sparse (e.g., hydrogen halides and glycine are
released from dust, De Keyser et al. 2017; Altwegg et al. 2016).
Incidentally, we note that, in presence of rapidly subliming (i.e.,
small and dirty) ice grains, the trend would have been opposite.
The VIRTIS-H observations suggest an increase by ∼ 20%
of the bolometric albedo in the -2 d to 21 d wrt perihelion pe-
riod, when the comet was the most active, possibly associated
with an increase of S heat. This trend would be in line with an
increased number of small particles at perihelion time, or alter-
natively, with enhanced degradation of dark material. The 67P
nucleus surface showed a global enhancement of water ice con-
tent near perihelion (Ciarniello et al. 2016; Fornasier et al. 2016).
The observed A increase would be in line with an expected in-
creased amount of icy grains in the inner coma of 67P. On the
other hand, this does not explain the trend observed for S heat.
h
law, close to the r−0.5
6. Summary and conclusion
Spectra of the dust 2 -- 4.5 µm continuum radiation have been ac-
quired with the VIRTIS-H experiment aboard the Rosetta mis-
sion to comet 67P. Through spectral fitting, we measured the
dust color temperature, bolometric albedo and 2 -- 2.5 µm color.
From the analysis of data acquired from 3 June to 29 Oct. 2015
(rh = 1.24 -- 1.55 AU) at line-of-sight tangent altitudes between
0.5 and 10 km, the following results were obtained:
-- At phase angles ∼90◦, the color temperature varied from 260
K to 320 K, and followed a r−0.6
vari-
ation expected from the balance between absorbed solar ra-
diation and radiated thermal energy. A 20% increase of the
bolometric albedo is observed near perihelion.
-- A mean dust color of 2%/100 nm and superheating factor
of 1.19 are measured at 90◦ phase angle, consistent with
previous VIRTIS-H measurements (Bockelée-Morvan et al.
2017).
-- A decrease of the color temperature with increasing phase
angle is observed, at a rate of ∼ 0.3 K/◦. It can be explained
by the presence of large porous particles, with low ther-
mal inertia, presenting a significant day-to-night temperature
contrast.
-- A large spectral phase reddening is measured. The value
(0.032 %/100 nm/◦) is higher then values measured for the
nucleus in the near IR (0.013 -- 0.018 %/100 nm/◦, Ciarniello
et al. 2015; Longobardo et al. 2017). This phase reddening
can be related to the roughness of the dust particles.
h
-- The bolometric albedo was found to increase from 0.05 to
0.14 (i.e., by a factor 2.5) with increasing tangent altitude
(so-called elevation in the paper) from 0.5 to 8 km. A de-
crease of the color temperature and color with increasing al-
titude is marginally observed. Possible explanations include
the presence in the inner coma of dark particles on ballistic
trajectories, and changes in particle composition.
-- Evidence for grain fragmentation, or disappearance of icy
grains, are not seen.
In future papers, we seek exploring the infrared continuum
images obtained with VIRTIS-H to obtain further constraints on
the dust coma of comet 67P.
Bockelée-Morvan et al.: 67P dust coma
Acknowledgements
The authors would like to thank the following institutions and
agencies, which supported this work: Italian Space Agency (ASI
- Italy), Centre National d'Etudes Spatiales (CNES -- France),
Deutsches Zentrum für Luft- und Raumfahrt (DLR -- Germany),
National Aeronautic and Space Administration (NASA -- USA).
VIRTIS was built by a consortium from Italy, France and Ger-
many, under the scientific responsibility of the Istituto di As-
trofisica e Planetologia Spaziali of INAF, Rome (IT), which lead
also the scientific operations. The VIRTIS instrument develop-
ment for ESA has been funded and managed by ASI, with con-
tributions from Observatoire de Meudon financed by CNES and
from DLR. The instrument industrial prime contractor was for-
mer Officine Galileo, now Leonardo company in Campi Bisen-
zio, Florence, IT. The authors wish to thank the Rosetta Science
Ground Segment and the Rosetta Mission Operations Centre for
their fantastic support throughout the early phases of the mis-
sion. The VIRTIS calibrated data shall be available through the
ESA's Planetary Science Archive (PSA) Web site. With fond
memories of Angioletta Coradini, conceiver of the VIRTIS in-
strument, our leader and friend. D.B.M. thanks E. Lellouch for
enlightening discussions.
References
Altobelli, N., Spilker, L. J., Leyrat, C., & Pilorz, S. 2008, Planet. Space Sci., 56,
Altwegg, K., Balsiger, H., Bar-Nun, A., et al. 2016, Science Advances, 2,
134
e1600285
Arakawa, S., Tanaka, H., Kataoka, A., & Nakamoto, T. 2017, A&A, 608, L7
Bardyn, A., Baklouti, D., Cottin, H., et al. 2017, MNRAS, 469, S712
Beck, P., Pommerol, A., Thomas, N., et al. 2012, Icarus, 218, 364
Bentley, M. S., Schmied, R., Mannel, T., et al. 2016, Nature, 537, 73
Bertini, I., La Forgia, F., Tubiana, C., et al. 2017, MNRAS, 469, S404
Bertini, I., Forgia, F. L., Fulle, M., et al. 2018, MNRAS, 482, 2924
Bockelée-Morvan, D., Debout, V., Erard, S., et al. 2015, A&A, 583, A6
Bockelée-Morvan, D., Crovisier, J., Erard, S., et al. 2016, MNRAS, 462, S170
Bockelée-Morvan, D., Rinaldi, G., Erard, S., et al. 2017, MNRAS, 469, S443
(Erratum 2017, MNRAS, 469, S842)
Boissier, J., Bockelée-Morvan, D., Biver, N., et al. 2012, A&A, 542, A73
Chamberlain, J. W., & Hunten, D. M. 1987, International Geophysics Series, 2nd
Ed. Academic Press, 36
Ciarniello, M., Capaccioni, F., Filacchione, G., et al. 2015, A&A, 583, A31
Ciarniello, M., Raponi, A., Capaccioni, F., et al. 2016, MNRAS, 462, S443
Colangeli, L., Lopez-Moreno, J. J., Palumbo, P., et al. 2007, Space Sci. Rev.,
Consolmagno, G. J., Schaefer, M. W., Schaefer, B. E., et al. 2013,
128, 803
Planet. Space Sci., 87, 146
Coradini, A., Capaccioni, F., Drossart, P., et al. 2007, Space Sci. Rev., 128, 529
De Keyser, J., Dhooghe, F., Altwegg, K., et al. 2017, MNRAS, 469, S695
Drossart, P., et al., 2000, SPIE, 4131, 78
Feller, C., Fornasier, S., Hasselmann, P. H., et al. 2016, MNRAS, 462, S287
Fornasier, S., Mottola, S., Keller, H. U., et al. 2016, Science, 354, 1566
Frattin, E., Cremonese, G., Simioni, E., et al. 2017, MNRAS, 469, S195
Fray, N., Bardyn, A., Cottin, H., et al. 2016, Nature, 538, 72
Fulle, M., Ivanovski, S. L., Bertini, I., et al. 2015, A&A, 583, A14
Fulle, M., Altobelli, N., Buratti, B., et al. 2016, MNRAS, 462, S2
Fulle, M., Della Corte, V., Rotundi, A., et al. 2017, MNRAS, 469, S45
Gehrz, R. D., & Ney, E. P. 1992, Icarus, 100, 162
Gerig, S.-B., Marschall, R., Thomas, N., et al. 2018, Icarus, 311, 1
Hammel, H. B., Telesco, C. M., Campins, H., et al. 1987, A&A, 187, 665
Hanner, M. 2003, J. Quant. Spec. Radiat. Transf., 79, 695
Harris, A. W. 1998, Icarus, 131, 291
Ivanovski, S. L., Zakharov, V. V., Della Corte, V., et al. 2017a, Icarus, 282, 333
Ivanovski, S. L., Della Corte, V., Rotundi, A., et al. 2017b, MNRAS, 469, S774
Keller, H. U., Barbieri, C., Lamy, P., et al. 2007, Space Sci. Rev., 128, 433
Keller, H. U., Mottola, S., Hviid, S. F., et al. 2017, MNRAS, 469, S357
Kissel, J., Altwegg, K., Clark, B. C., et al. 2007, Space Sci. Rev., 128, 823
Kolokolova, L., Hanner, M. S., Levasseur-Regourd, A.-C., & Gustafson, B. Å. S.
2004, Comets II, 577
Kurucz R. L., 1992, in Rabin D. M., Jefferies, J. T., eds, Proc. IAU Symp. Vol.
154, Synthetic Infrared Spectra in Infrared Solar Physics. Kluwer, Dordrecht
Langevin, Y., Hilchenbach, M., Ligier, N., et al. 2016, Icarus, 271, 76
Le Gall, A., Leyrat, C., Janssen, M. A., et al. 2014, Icarus, 241, 221
Leinert, C., Richter, I., Pitz, E., & Planck, B. 1981, A&A, 103, 177
Leyrat, C., Ferrari, C., Charnoz, S., et al. 2008, Icarus, 196, 625
Longobardo, A., Palomba, E., Capaccioni, F., et al. 2017, MNRAS, 469, S346
Markkanen J., Agarwal J., Väisänen T., Penttilä A., Muinonen K., 2018, ApJ,
Macke, R. J., Consolmagno, G. J., & Britt, D. T. 2011, Meteoritics and Planetary
868, L16
Science, 46, 1842
Mannel T., Bentley M. S., Schmied R., Jeszenszky H., Levasseur-Regourd A. C.,
Romstedt J., Torkar K., 2016, MNRAS, 462, S304
Mannel, T., Bentley, M. S., Boakes, P., et al. 2017, European Planetary Science
Mason C. G., Gehrz R. D., Jones T. J., Woodward C. E., Hanner M. S., Williams
Congress, 11, EPSC2017-258
D. M., 2001, ApJ, 549, 635
Merouane, S. et al., 2016, A&A, 596, A87
Moreno, F., Muñoz, O., Gutiérrez, P. J., et al. 2017, MNRAS, 469, S186
Moreno, F. et al. 2018, AJ, 156, 237
Opeil, C. P., Consolmagno, G. J., & Britt, D. T. 2010, Icarus, 208, 449
Ott, T., Drolshagen, E., Koschny, D., et al. 2017, MNRAS, 469, S276
Reach, W. T., Vaubaillon, J., Lisse, C. M., Holloway, M., & Rho, J. 2010, Icarus,
208, 276
Riedler, W., Torkar, K., Jeszenszky, H., et al. 2007, Space Sci. Rev., 128, 869
Rinaldi, G., Fink, U., Doose, L., et al. 2016, MNRAS, 462, S547
Rinaldi, G., Della Corte, V., Fulle, M., et al. 2017, MNRAS, 469, S598
Schiffer, R. 1985, A&A, 148, 347
Schloerb, F. P., Gulkis, S., Biver, N., et al. 2017, AAS/Division for Planetary
Sciences Meeting Abstracts #49, 49, 415.06
Schröder, S. E., Grynko, Y., Pommerol, A., et al. 2014, Icarus, 239, 201
Shou, Y., Combi, M., Toth, G., et al. 2017, ApJ, 850, 72
Spencer, J. R., Lebofsky, L. A., & Sykes, M. V. 1989, Icarus, 78, 337
Telesco, C. M., Decher, R., Baugher, C., et al. 1986, ApJ, 310, L61
Tokunaga A. T., Golisch W. F., Griep D. M., Kaminski C. D., & Hanner M. S.
1986, AJ, 92, 1183
Vincent, J.-B., A'Hearn, M. F., Lin, Z.-Y., et al. 2016, MNRAS, 462, S184
Appendix A:
Table A.1. Log of VIRTIS-H observations and retrieved dust spectral
and temperature properties. (This Table is not available in the present
version.)
Appendix B: Thermal model for dust particles
A simple approach, inherited from asteroid studies, is used to
model the particle surface temperature as a function of local
time. Spencer et al. (1989) showed that, for smooth objects with
the Sun in the equatorial plane, the diurnal temperature profile
can be parameterized by a quantity Θ called thermal parameter.
Θ is function of the thermal inertia Γ, angular rotation rate ω,
subsolar equilibrium temperature TSS, emissivity , and Stefan-
Boltzmann constant σ:
√
Γ
ω
σT 3
SS
,
Θ =
with
(cid:113)
(B.1)
(B.2)
KC∗
p,
Γ =
where K is the thermal conductivity and C∗
heat capacity.
p is the volumetric
The subsolar equilibrium temperature for dark and smooth
objects is given by:
TSS =
394
r0.5
h ()0.25
[K].
(B.3)
Article number, page 11 of 13
A&A proofs: manuscript no. dustprop-VR-accepted
Spencer et al. (1989) provide diurnal temperature profiles
on the equator as a funtion of Θ. In their Fig. 2, they show the
variation of TMAX/TSS, TMIN/TSS with Θ, where TMAX and TMIN
are respectively the maximum and minimum surface tempera-
tures, and TSS the subsolar temperature for a non-rotating body
(Eq. B.3). We described the surface temperature at latitude θ(cid:48) --
π/2 and longitude φ(cid:48) (subsolar point at equator, i.e., θ(cid:48) = 90◦ and
φ(cid:48) = 0◦) by:
(cid:16)
Tday(θ
max
(cid:48)
, Θ) =
(cid:48)
, φ
TMAX(Θ)(sin(θ
(cid:48)) cos(φ
(cid:48)))0.25, Tnight(θ
(cid:48)
(cid:48)
, φ
, Θ)
(cid:17)
(B.7)
(cid:48)
(cid:48)
, φ
(cid:48))0.25,
, Θ) = TMIN(Θ) sin(θ
Tnight(θ
(B.8)
where θ(cid:48) is in the range (0, π), and φ(cid:48) is in the range ( -- π, π).
This (θ(cid:48), φ(cid:48)) dependence corresponds to NEATM for Θ = 0, and
should approximate well the temperature distribution when Θ >
0, except that the shift of the maximum temperature in the af-
ternoon for nonzero Θ is not represented. TMAX(Θ) and TMIN(Θ)
were taken from Spencer et al. (1989). Figure 9 shows diurnal
temperature curves for different Θ.
We then computed the 3 -- 5 µm thermal emission as a func-
tion of phase angle, considering the surface elements facing the
observer (cf. NEATM model from Harris (1998)). We used a bi-
modal distribution for the grains, consisting of isothermal grains
at T =Tiso (e.g., rapidly spinning dust particles, or grains with
size smaller than δtherm), and slowly spinning/low thermal inertia
particles, with a temperature profile described by Θ (Eq B.7 --
B.8). This approach follows the one adopted by Leyrat et al.
(2008) to explain the phase variation of the color temperature of
Saturn's C ring. The relative contribution of isothermal particles
is given by fiso, which determines the optical depth contribution
of isothermal particles ( fiso = τiso/(τiso + τnon−iso)). The optical
depth is proportional to the integral over size range of the size
distribution times particle cross-section (e.g. Leyrat et al. 2008).
The temperature of the isothermal particles is expressed as:
Tiso =
fheatTequ
0.25
,
(B.9)
where Tequ, which corresponds to the equilibrium temperature
for an emissivity equal to 1, is given in Eq. 1. The parameter fheat
allows us to investigate dust particles heated above the equilib-
rium temperature. This temperature excess is expected for small
particles made of absorbing material (Kolokolova et al. 2004,
and references therein).
Table B.2 summarizes the free parameters of the model for
computing synthesized spectra of dust thermal emission. By fit-
ting a blackbody function to these spectra, the superheating fac-
tor as a function of phase angle can be derived and compared to
VIRTIS-H measurements.
Objects with higher Θ will have lower day-to-night temper-
ature contrast, lower subsolar temperature and maximum tem-
perature in the afternoon. Objects with low thermal inertia and
low rotation rate have low Θ, so high temperature contrasts with
maximum temperature near midday.
The diurnal skin depth is given by (Le Gall et al. 2014):
(cid:114)
δtherm =
Γ
C∗
p
2
ω
.
(B.4)
The surface becomes isothermal with time of the day and depth
for Θ > 50 (Spencer et al. 1989). However, diurnal tempera-
ture profiles parameterized by Θ (Spencer et al. 1989) are in-
valid for objects with sizes in the range or lower than the diurnal
skin depth δtherm. Due to solar heat penetration, they should be
isothermal. We computed Θ and δtherm for a range of parameters
appropriate to 67P porous dust aggregates (Table B.1). The ther-
mal conductivity Kmat, specific heat capacity Cp, and volumet-
ric density ρmat of the material composing the monomers (i.e.,
subunits) correspond to values measured for carbonaceous mete-
orites. The adopted ρmat value is also consistent with estimations
of the compacted bulk density of 67P dust particles (Fulle et al.
2017). The volumetric heat capacity and thermal conductivity of
the aggregates depend on their porosity φ. We have:
p = Cp(1 − φ)ρmat.
C∗
For the thermal conductivity, we used the empirical formula
derived by Arakawa et al. (2017) for highly porous dust aggre-
gates, giving:
(B.5)
K = ckKmat(1 − φ)2.
(B.6)
The coefficient of proportionality ck is given in Arakawa et
al. (2017). Comparing with laboratory measurements, Arakawa
et al. (2017) found that this porosity dependence can be used for
φ > 0.5. The thermal conductivity of aggregates derived from
this empirical formula is K = 0.00714 to 0.000285 W m−1 K−1
for φ = 0.5 to 0.9.
Rotation frequencies (ω/2π) of 67P grains have been es-
timated by dust dynamical simulations (Fulle et al. 2015;
Ivanovski et al. 2017a,b), and also measured using the light
curves of individual dust particles (Fulle et al. 2015). Computed
values for > 100 µm size particles range from 0.03 -- 4 Hz depend-
ing of grain characteristics (shape, size and density), gas flow en-
vironment and initial conditions (Fulle et al. 2015; Ivanovski et
al. 2017a). Measurements in the low activity phase of the comet
show that the most probable rotation frequency of the grains is
below 0.15 Hz (Fulle et al. 2015). The simulations show that par-
ticles experience oscillations before acquiring a full rotation at a
few kilometers from the surface, with the rotation axis perpen-
dicular to the flow direction. Modeling the surface temperature
of oscillating grains is beyond the scope of this paper.
The thermal parameter, diurnal skin depth and thermal iner-
tia for porosities of 0.5 -- 0.9 and rotation frequencies of 0.0001 --
0.1 Hz are shown in Fig. B.1. Calculations were performed with
the parameters given in Table B.1. Thermal inertias are low (8 --
80 J m−2 K−1 s−1/2). The diurnal skin depth ranges from ∼ 0.01
to 0.5 cm, and the thermal parameter varies from less than 1
(strong day-to-night contrast) for slowly rotating grains to val-
ues ∼ 30 for relatively compact aggregates (φ = 0.5) with rota-
tion frequency of 0.1 Hz.
Article number, page 12 of 13
Table B.1. Model fixed parameters.
Bockelée-Morvan et al.: 67P dust coma
Value Note
CM chondrites
Parameter
Aggregate subunits:
Thermal conductivity Kmat
(W/m/K)
Heat capacity
(J/kg/K)
Density
(kg/m3)
Aggregate
Emissivity
a Opeil et al. (2010) b Consolmagno et al. (2013) c Macke et al. (2011)
300 K -- Chondrites
2420c Orgueil
0.9
0.7a
700b
Cp
ρmat
Table B.2. Model free parameters.
Parameter Description
Phase angle
α
Thermal parameter
Θ
fiso
Optical depth contribution
of isothermal particles
Factor to equilibrium temperature
of isothermal particles
fheat
Fig. B.1. Thermal parameter Θ, diurnal skin depth δtherm and thermal
inertia Γ as a function of particle rotation frequency. The different curves
correspond to different porosities, with values indicated on the plots.
Article number, page 13 of 13
|
1605.06291 | 1 | 1605 | 2016-05-20T11:13:43 | Dust Impact Monitor (SESAME-DIM) on board Rosetta/Philae: Millimetric particle flux at comet 67P/Churyumov-Gerasimenko | [
"astro-ph.EP"
] | The Philae lander of the Rosetta mission, aimed at the in situ investigation of comet 67P/C-G, was deployed to the surface of the comet nucleus on 12 Nov 2014 at 2.99 AU heliocentric distance. The Dust Impact Monitor (DIM) as part of the Surface Electric Sounding and Acoustic Monitoring Experiment (SESAME) on the lander employed piezoelectric detectors to detect the submillimetre- and millimetre-sized dust and ice particles emitted from the nucleus. We determine the upper limit of the ambient flux of particles in the measurement range of DIM based on the measurements performed with the instrument during Philae's descent to its nominal landing site Agilkia at distances of about 22 km, 18 km, and 5 km from the nucleus barycentre and at the final landing site Abydos. The geometric factor of the DIM sensor is calculated assuming an isotropic ambient flux of the submillimetre- and millimetre-sized particles. For the measurement intervals when no particles were detected the maximum true impact rate was calculated by assuming Poisson distribution of the impacts, and it was given as the detection limit at a 95% confidence level. Based on measurements performed with DIM, the upper limit of the flux of particles in the measurement range of the instrument was of the order of $10^{-8}-10^{-7}m^{-2}s^{-1}sr^{-1}$ during descent and $1.6\cdot10^{-9}m^{-2}s^{-1}sr^{-1}$ at Abydos on 13 and 14 Nov 2014. Considering particle speeds below escape velocity, the upper limit for the volume density of particles in the measurement range of DIM was constrained to $10^{-11}m^{-3}-10^{-12}m^{-3}$. Results of the calculations performed with the GIPSI tool on the expected particle fluxes during the descent of Philae were compatible with the non-detection of compact particles by the DIM instrument. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. AA_postprint
September 13, 2018
c(cid:13)ESO 2018
Dust Impact Monitor (SESAME-DIM) on board
Rosetta/Philae: Millimetric particle flux at comet
67P/Churyumov-Gerasimenko
Attila Hirn1,2,(cid:63), Thomas Albin3,4,2, István Apáthy1, Vincenzo Della Corte5, Hans-Herbert Fischer6, Alberto
Flandes7,2, Alexander Loose2, Attila Péter1, Klaus J. Seidensticker8, and Harald Krüger2
Hungary
D.F.
Germany
6
1
0
2
y
a
M
0
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
9
2
6
0
.
5
0
6
1
:
v
i
X
r
a
1 Centre for Energy Research, Hungarian Academy of Sciences, Konkoly Thege Miklós út 29-33, 1121 Budapest,
2 Max-Planck-Institut für Sonnensystemforschung, Justus-von-Liebig-Weg 3, 37077 Göttingen, Germany
3 Institut für Raumfahrtsysteme, University Stuttgart, Pfaffenwaldring 29, 70569 Stuttgart, Germany
4 Medical Radiation Physics, Faculty VI, Carl von Ossietzky University, Georgstrasse 12, 26121 Oldenburg, Germany
5 Institute for Space Astrophysics and Planetology (IAPS), National Institute for AstroPhysics (INAF), Via Fosso del
Cavaliere 100, 00133 Roma, Italy
Köln, Germany
6 Deutsches Zentrum für Luft- und Raumfahrt, Raumflugbetrieb und Astronautentraining, MUSC, Linder Höhe, 51147
7 Ciencias Espaciales, Instituto de Geofísica, Universidad Nacional Autónoma de México, Coyoacán 04510, México,
8 Deutsches Zentrum für Luft- und Raumfahrt, Institut für Planetenforschung, Rutherfordstrasse 2, 12489 Berlin,
Received; accepted
ABSTRACT
Context. The Philae lander of the Rosetta mission, aimed at the in situ investigation of comet 67P/Churyumov-
Gerasimenko, was deployed to the surface of the comet nucleus on 12 November 2014 at 2.99 AU heliocentric dis-
tance. The Dust Impact Monitor (DIM) as part of the Surface Electric Sounding and Acoustic Monitoring Experiment
(SESAME) on the lander employed piezoelectric detectors to detect the submillimetre- and millimetre-sized dust and
ice particles emitted from the nucleus.
Aims. We determine the upper limit of the ambient flux of particles in the measurement range of DIM based on the
measurements performed with the instrument during Philae's descent to its nominal landing site Agilkia at distances
of about 22 km, 18 km, and 5 km from the nucleus barycentre and at the final landing site Abydos.
Methods. The geometric factor of the DIM sensor was calculated assuming an isotropic ambient flux of the submillimetre-
and millimetre-sized particles. For the measurement intervals when no particles were detected the maximum true impact
rate was calculated by assuming Poisson distribution of the impacts, and it was given as the detection limit at a
95% confidence level. The shading by the comet environment at Abydos was estimated by simulating the pattern of
illumination on Philae and consequently the topography around the lander.
Results. Based on measurements performed with DIM, the upper limit of the flux of particles in the measurement range
of the instrument was of the order of 10−8 − 10−7m−2s−1sr−1 during descent. The upper limit of the ambient flux of
the submillimetre- and millimetre-sized dust and ice particles at Abydos was estimated to be 1.6 · 10−9m−2s−1sr−1 on
13 and 14 November 2014. A correction factor of roughly 1/3 for the field of view of the sensors was calculated based
on an analysis of the pattern of illumination on Philae.
Conclusions. Considering particle speeds below escape velocity, the upper limit for the volume density of particles in
the measurement range of DIM was constrained to 10−11 m−3 − 10−12 m−3. Results of the calculations performed with
the GIPSI tool on the expected particle fluxes during the descent of Philae were compatible with the non-detection of
compact particles by the DIM instrument.
Key words. Rosetta -- Philae -- comets -- cometary dust -- piezoelectric detectors -- 67P/Churyumov-Gerasimenko
1. Introduction
to
comet
After
its more
67P/Churyumov-Gerasimenko
67P/C-G),
the Rosetta spacecraft (Glassmeier et al. 2007) reached
its target on the 6 August 2014 to start a series of in
situ measurements from around the nucleus. In addition
cruise
(hereafter
than
10-year
(cid:63) contact address: e-mail: [email protected]
to the 11 orbiter experiments, Rosetta also carried a
lander, named Philae (Bibring et al. 2007), which was
deployed onto the surface of the nucleus of the comet on
12 November 2014 (Biele et al. 2015).
The Dust Impact Monitor (DIM) of the Surface Electric
Sounding and Acoustic Monitoring Experiment (SESAME)
package (Seidensticker et al. 2007) on board the lander was
one of the instruments that were active and operating not
Article number, page 1 of 17
A&A proofs: manuscript no. AA_postprint
only during the first few days after the landing of Philae
(first science sequence -- FSS) on the nucleus surface at
the final landing site Abydos, but also during the sepa-
ration, descent, and landing (SDL) phase of the mission.
The DIM instrument was designed to measure the flux of
submillimetre- and millimetre-sized dust and ice particles
emitted from the nucleus by means of 3 × 3 piezoelectric
sensor segments made of PNZT7700 (Pb, Ni, Zi, Ti, here-
after referred to as PZT) and mounted on three sides of
a cube. From the signal properties measured with the as-
sociated SESAME common electronics, the mass and the
speed of the impacting particles could be constrained as-
suming given density and elastic material properties. Since
the sensor sides are oriented in three mutually orthogo-
nal directions, an assessment of the directionality of the
impacting particles might be also made provided that the
number of impacts is statistically sufficient (Seidensticker
et al. 2007; Flandes et al. 2013, 2014).
DIM was operated during three mission phases of Philae
at the comet (Krüger et al. 2015): before separation, during
descent, and at the final landing site. In the mission phase
before Philae's separation from Rosetta, at altitudes be-
tween approximately 8 and 23 km from the nucleus surface,
DIM was significantly obscured by structures of Rosetta
and no particles were detected. During Philae's descent to
its nominal landing site Agilkia, DIM detected one approx-
imately millimetre-sized particle at an altitude of 2.4 km.
This is the closest ever detection at a cometary nucleus by a
dedicated in situ dust detector. The material properties of
the detected particle are compatible with a porous particle
having a bulk density of approximately 250 kg m−3. At Phi-
lae's final landing site, Abydos, DIM detected no particle
impacts.
In this paper we present upper limits of the flux of parti-
cles in the measurement range of the DIM instrument in the
two operational phases after the release of Philae. Phases
preceding the release are not considered in our analysis be-
cause of the complexity of the shielding geometry and a
reduced geometric factor of the shaded sensors. Measure-
ments of the particle flux on the orbiter are provided by the
Grain Impact Analyser and Dust Accumulator (GIADA)
team for this period with much greater sensitivity and a
different measurement range (Della Corte et al. 2015; Ro-
tundi et al. 2015). We discuss in detail the effects of shading
by the detector frame and the body of the lander on the
geometric factor of the DIM sensor, and also address the
effects of the local environent. A rough estimation on the
upper limit of the volume density of particles in DIM's mea-
surement range is presented.
2. Dust Impact Monitor
2.1. Detector geometry
The DIM cube of dimensions 71.5 mm× 71.5 mm× 69.0 mm
is mounted on the top face of the lander, above Philae's
balcony, with sensor sides pointing in the +X, +Y, and +Z
directions in the Philae coordinate system (Fig. 1). The -X
and -Y sides are covered with aluminum plates, whereas the
-Z side is left open for cabling and mounting purposes. The
three PZT segments on the active sides have dimensions
50.0 mm × 16.2 mm × 1.0 mm and they are separated by
1.5 mm (Fig. 2). They all lie 2.3 mm below the frame of the
DIM cube. Impacts incident on different sensor segments
Article number, page 2 of 17
belonging to a given sensor side are not distinguished by
the electronics.
The PZT segments are significantly shaded by Philae's
structure and by the sensor frame; the amount of shad-
ing differs from one side of the sensor to the other (see
Fig. 1). The field of view (FoV) of the +X side is limited
mostly to the +Z direction owing to the structure of the
lander. However, the +Y sensor side, being closer to the
edge of the structure, is only partially shielded, mostly by
Solar Array Panel 1, for particles approaching from the -- Z
half-space. Because it is close to the drill box of the drill,
sample, and distribution (SD2) subsystem protuding ap-
proximately 150 mm above Philae's solar hood and 75 mm
above DIM's Z side (see Fig. 1), the +Z sensor side is also
partially shielded from particles coming from the (-X; -Y;
+Z) region, but that side is still the least shielded of the
three active sides of DIM.
2.2. Measurement technique
When a PZT segment on DIM is hit by a particle, the sen-
sor generates a damped, closely sinusoidal electrical signal.
Calibration measurements performed on the ground with
different types of test particles impacting on the DIM sensor
at different speeds (Péter 2002; Flandes et al. 2013, 2014)
have shown that the impacts can be described and analysed
using the Hertz theory of contact mechanics (Hertz 1882;
Johnson 2003). After recording the amplitude (Um) and the
width of the first half-sine pulse (Tc), it is possible to con-
strain the radius R and the impact speed v of the particle
(Seidensticker et al. 2007; Flandes et al. 2013) according to
Eqs. 1 and 2,
Um =
3.03d33E0.4
r ρ0.6R2v1.2
C
(cid:33)1/5
,
,
(1)
(2)
(cid:32)
R5ρ2
vE2
r
Tc = 5.09
where d33 is the piezoelectric stress constant of the PZT,
Er the combined reduced Young's modulus of the sensor
and the impinging particle, ρ the density of the impacting
particle, and C the capacitance of the sensor plate.
The signals coming from the sensor segments are first
amplified with a logarithmic amplifier. The instrument de-
tects a signal only if the amplified signal crosses an adap-
tive detection threshold voltage defined as the sum of an
adjustable margin and the signal average determined by
the DIM electronics with a time constant of approximately
1 s. The margin can be increased in steps of 10 dB in the
range from 10 dB to 70 dB. Each step changes the thresh-
old voltage by approximately 0.3 V. If the amplified signal
crosses the detection threshold less than 1 ms after the sin-
gle event measurement is initialized, the signal is classified
as a "false event", else it is accepted as the beginning of
a potential real impact. If no second threshold crossing is
detected within a time interval pre-defined for the given op-
erational mode, the event is ruled out as a "long event". A
new single event measurement is initialized only after some
waiting and latency periods adding up to a total instrument
dead time of approximately 10 ms. A more detailed descrip-
tion of the DIM signal processing is given in Péter & Lucás
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
(2001); Fischer (2014) and Krüger et al. (2015). Flandes
et al. (2013) have shown that signals with amplitudes in the
interval 0.2 mV < Um < 15 mV deliver measurement val-
ues within the expected theoretical behaviour and from this
they have determined an approximated experimental range
of detection radii based on Eq. 1 (Krüger et al. 2015). The
intervals are reported for the different operational modes in
Sects. 2.3.1 and 2.3.2.
2.3. Operation
After Rosetta's launch in March 2004, the health and the
performance of the DIM instrument were regularly checked
and interference tests were executed in the frame of payload
check-outs performed approximately every six months until
the spacecraft entered deep space hibernation in 2011. After
the wake-up of Rosetta and its lander Philae in 2014, fur-
ther tests were performed in order to guarantee that DIM
was working properly. A detailed description of these op-
erations (health-checks, tests, and measurements) is given
in Fischer (2012) and Krüger et al. (2015) and the cor-
responding SESAME Lander Instrument Operations Re-
quests (LIOR) documents. In the present paper we focus
exclusively on the measurement modes used in the SDL
and FSS phases of the mission.
2.3.1. Measurement mode during the separation, descent,
and landing phase
During the SDL phase, measurements were performed in
the so-called Burst Continuous Test2 mode (BCT2). This
measurement mode delivers the measured raw peak am-
plitude Um, the impact duration Tc, and the time of de-
tection for the first 350 detected events on a given sensor
side. The total number of detected events, false events, and
long events are also recorded. On 12 November 2014, after
Philae's separation from Rosetta at 08:35:00 UTC, three
measurement blocks were conducted at distances of about
22 km, 18 km, and 5 km from the nucleus barycentre. In
each block all three sensor sides were operated. Measure-
ment times were 100 s or 200 s.
During the tests of the descent with the Philae ground
reference model performed at the Deutsches Zentrum für
Luft und Raumfahrt (DLR) in Cologne, a cross-talk with
the Maximum Power Point Tracking (MPPT) electronics of
the solar arrays was identified, which resulted in a high rate
of false signals interpreted as detected events by the DIM
electronics. For the most part, these events were recorded
for only a few seconds at the beginning of the blocks. The
same behaviour could be observed in the flight data, which
means that those measurements in which the number of
false signals exceeded 350 could not be used for detecting
particle impacts.
The detection intervals in terms of particle radius
for BCT2 measurements during Philae's descent were
[0.5 mm − 6.5 mm] and [0.9 mm − 6.5 mm] for margin levels
of 40 dB and 50 dB, respectively, used during these mea-
surement blocks (Krüger et al. 2015).
2.3.2. Measurement mode during the first science sequence
phase
In the measurement blocks of the FSS DIM was operated
in Burst Continuous (BC) mode. The BC mode delivers the
counts for impacts with a particular [Um, Tc] combination.
The Um and Tc values are stored in a compressed way in
memory cells of different sizes, depending on the expected
frequency of such events. The event times are not registered
by the instrument.
Each BC mode measurement lasted for 557−558 s; mar-
gins were set either to 30 dB with radius detection interval
[0.25 mm−1.5 mm] or 40 dB with [0.5 mm−1.5 mm](Krüger
et al. 2015).
3. Methods
3.1. Maximum impact rates
Provided that only one real impact was registered by DIM
during the scientific measurements in the SDL and the FSS
phases, it is reasonable to assume an isotropic distribution
of the particle trajectories. Moreover, we can suppose that
the impacts on the DIM sensor are independent events,
hence we can also assume that their occurrence follows a
Poisson distribution. For the periods when no detection was
made, we seek the detection limit, i.e. the value of the pa-
rameter λ of the Poisson distribution for which there is an
arbitrarily chosen 95% (2σ) probability that the number of
detected events N will exceed zero in a single measurement
P (N > 0) = 1 − P (N = 0) = 1 − λ0 exp(−λ)
= 1 − exp(−λ) = 0.95 ,
(3)
=
(0)!
thus
λ = − ln(0.05) ≈ 3 .
(4)
For the measurement block, when exactly one real sig-
nal was detected, again, only the upper limit of the ambient
flux can be estimated. We can define the upper limit of the
expected number of impacts as the highest value of λ for
which there is a 5% probability that the number of the de-
tected events N will be less than 2 in a single measurement:
P (N < 2) = P (N = 0) + P (N = 1) =
= (1 + λ) exp(−λ) = 0.05 ,
resulting in
λ ≈ 4.74 .
(5)
(6)
3.2. Geometric factor of the DIM sensor
The relation between the measured impact rates (N in s−1)
and the particle flux (Φ in m−2s−1sr−1) in the measurement
range of the sensor is given by
N = GΦ ,
(7)
where G is the geometric factor of the detector in m2sr.
Article number, page 3 of 17
A&A proofs: manuscript no. AA_postprint
3.2.1. Stand-alone PZT segment
First we consider only one single PZT segment in the XY
plane. The geometric factor is given as the sum of the ef-
fective areas seen from different directions. For an isotropic
particle flux coming from the Z > 0 half-space, G0 = πA,
where A = W L is the surface area of one PZT segment
(Fig. 2).
3.2.2. Shading effect of the frame
The outer structure of the DIM sensor that frames the PZTs
produces significant shading for particles incident to the
sensitive surface under highly oblique angles. The surface
of the sensor segments lies 2.3 mm below the outer surface of
the frame, which is significant compared to the dimensions
of the segments (W = 16 mm; L = 50 mm). For example, if
particles come along a direction for which φ = 0◦ and θ >
87◦, or likewise, for which φ = 90◦ and θ > 82◦ the sensor
frame completely prevents the particles from reaching the
PZT (see Fig. 2).
The geometric factor, in this case, can be calculated an-
alytically and it is identical to the geometric factor of a
radiation detector having rectangular telescope geometry.
Thomas & Willis (1972) derived an analytical formula for
the general case of two rectangular areas having sides of
length 2X1, 2Y1 and 2X2, 2Y2 with Z being their sepa-
ration. The geometry of one DIM PZT segment with the
sensor frame around it corresponds to the special case of
equal rectangular areas with dimensions 2X1 = 2X2 = L
and 2Y1 = 2Y2 = W ; the separation between the two areas
is identical to the depth of the frame Z = d. The geometric
factor Gf then can be expressed as
√
Gf = 2L
√
d2 + W 2 arctan
L√
d2 + W 2
− 2Ld arctan
L
d
+
+2W
d2 + L2 arctan
W√
d2 + L2
− 2W d arctan
W
d
+
+d2 ln
(d2 + W 2)(d2 + L2)
(d2 + W 2 + L2)d2 .
(8)
If we calculate the limit of Gf as d tends to 0, we get
Gf = G0.
3.2.3. Shading by the structure of Philae
In order to consider the shading effects of Philae's struc-
ture and the other payloads on the DIM's FoV, numerical
simulations were performed with a virtual isotropic particle
flux because -- owing to the complexity of the structure --
the problem could not be solved analytically. The DIM sen-
sor was simulated with its frame mount (see Sect. 3.2.2),
whereas the CAD model of Philae (Bernd Chares, priv.
communication) was slightly simplified to reduce computa-
tional time, e.g. we neglected the lander feet and reduced
the shape of the SD2 drill tower to a properly sized cuboid.
The DIM PZT segments were divided into 800 identical
squares with surface area of 1 mm2 each. On each square,
32,400 particles were generated with an isotropic flux from
a half-space, which means a total number of approximately
Article number, page 4 of 17
78 million particles simulated per DIM side. For each lin-
early propagating particle the simulation checks if the trail
is within DIM's FoV or intersected by Philae's structures.
The ratio of actual impacting vs. the total number of sim-
ulated particles is named the detection ratio.
3.2.4. Calculated geometric factors
The detection ratio for each DIM side and the values of
the geometric factor calculated for the cases described in
Sect. 3.2 are summarized in Table 1. The geometric factors
are reduced by 17% for all three sides if only the shield-
ing effect of the sensor frame is taken into account (inher-
ent shading), whereas if the shielding by the structure and
payloads of the lander are also considered the values are re-
duced by 56%, 37%, and 33% respectively for sensor sides
X, Y, and Z.
The detection ratio decreases to the corners and the
edges for each segment owing to the shading by the DIM
frame. The least shielding occurs for the Z-side of DIM.
This side is only slightly shielded by the SD2 drill tower
and the DIM mounting frames. The analytical calculations
performed in Sect. 3.2.2 also served for verification of the
numerical model used in Sect. 3.2.3.
4. Results
4.1. Estimation of the maximum flux during descent
Estimations for the upper limit of the flux during descent
were performed only for those measurements for which the
effective measurement time was higher than 0 s. The re-
sults are summarized in Table 2. The sequential number of
the measurements in Col. 1 are according to Krüger et al.
(2015). Column 9 corresponds only to the number of real
detections; false signals are not included. The maximum
impact rate (Nmax.) in Col. 10 was defined as the ratio of
parameter λ of the Poisson distribution calculated accord-
ing to Eqs. (4) and (6) respectively for the non-detection
and the single detection of particle impacts to the effective
measurement time shown in Col. 6. Finally, Φmax. in Col. 11
was calculated according to Eq. (7).
4.2. Estimation of the maximum flux at Abydos
Since no impact signals were detected at all at the final
landing site Abydos during the FSS, the measurement times
for each sensor side were summed:
5(cid:88)
2(cid:88)
Tmeas., i, total =
Tmeas., i, j, k,
(9)
j=1
k=1
where Tmeas., i, j, k is the measurement time for sensor
side i in the kth measurement of FSS DIM Block number
j. The maximum impact rate for the ith sensor side in the
above case is
Nimp., i, max. = λmax./Tmeas., i, total ,
(10)
where λmax. is calculated according to Eq. 4.
The upper limits of the flux values (Φmax., see Table 3)
for particles in the measurement range of the DIM sensor
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
were estimated using Eq. 7 and by considering the geomet-
ric factors calculated for the lander configuration (see values
in Table 1).
4.3. Particle flux and the topography of Abydos
The topography at Abydos is a major concern with regard
to the potential flux detection, in addition to the activity
of the comet and the self shading by the structure of Phi-
lae or the sensor itself. The images of the Çiva cameras
(Bibring et al. 2015) at Abydos show that the lander is
partially surrounded by an obstacle that prevents the sun-
light from reaching the solar panels for long periods of time.
Right after landing, in November 2014, the panels received
sunlight for less than two hours per comet day, which was
not enough to charge the lander's batteries. Because under-
standing the illumination conditions on the lander would
help us to determine the topography at Abydos, Flandes et
al. (2016; document in preparation) simulated the pattern
of illumination on Philae assuming that the lander is par-
tially surrounded by a sinusoid-shaped barrier with a height
equivalent to approximately three times the height of the
lander (see Fig. 3). A top view of this configuration would
set the lander in a trough or peak of this sinusoid. In this
simulation the Sun always moves along the horizon for the
lander (at low elevations < 22.5◦). Under these conditions,
the lander is illuminated for 1.5 hours out of the 12.4 hour
rotation period of the comet. The panel that receives the
largest amount of light is Panel 1 (which is parallel with the
Y side of DIM), followed by Panel 2. Panel 6 (top panel,
which is parallel with the Z side of DIM) and Panel 3 receive
very little sunlight.
If this geometry is accurate, the incoming particle flux
for DIM will be very much reduced as well (assuming that
neither the barrier nor the floor of Abydos in the immedi-
ate vicinity of the spacecraft are dust sources). The tem-
perature measured by the Multi-Purpose Sensors for Sur-
face and Subsurface Science (MUPUS) experiment at 90 --
130 K is well below the water sublimation temperature,
which implies that the immediate environment was not ac-
tive (Spohn et al. 2015). For practical purposes, the lander
could be considered to be inside a hypothetical partially
opened box where the floor of the site is its bottom and the
barrier/wall represents only three of its sides. One conclu-
sion is that particles could reach the sensor only through
the top of this box or from one of its sides. With this fur-
ther simplification, the general field of view of the sensor (if
considered at the centre of this box) would be reduced to
1/3 of the total, i.e. 4/3π sr. Given that the Sun never goes
far above the horizon in this simulation, the three most im-
portant parameters are the separation between the sides,
the depth of this box, and the maximum elevation of the
Sun. A variation of > 10% in any of these three parameters
would produce a similar variation in the illumination pat-
tern on the lander. Still, the geometric factor of the sensor
segments is already reduced to 33% -- 56% as a result of shad-
ing by the lander structure and some other payloads (see
Table 1), which means that further corrections are unnec-
essary as the uncertainties in other factors are significantly
higher.
Auster et al. (2015) estimated the local gravity as
g = 10−4 m s−2 at the Abydos landing site at a distance
of 2332 m from the barycenter of the nucleus. This gives
an escape velocity of approximately 0.7 m s−1. If particle
speeds v below the escape velocity are considered, the ex-
pected speed range for particles in the measurement range
of DIM can be constrained to 0.1 m s−1−0.7 m s−1. An up-
per boundary for the volume density (nmax. = Φmax.Ωeff./v,
where Ωeff. is the effective solid angle) of such particles is
then of the order of 10−11 m−3 − 10−12 m−3.
5. Discussion
Although DIM was taking measurements starting from the
onset of comet activity of 67P/C-G in November 2014 at
about 3 AU heliocentric distance, only one single detection
of particle impact was made. During descent the relatively
short measurement times did not permit the measurement
of particles with good statistics. The total measurement
time was only 35 minutes, which -- owing to the large num-
ber of false signals resulting from the cross-talk from the
MPPTs -- was further reduced to less than 20 minutes noise-
free time. Hence, instead of the flux, the upper limit of the
flux of particles in the measurement range of DIM was de-
termined (see Table 2).
At the final landing site Abydos, the total measurement
time was an order of magnitude higher than for the descent,
but -- probably due to the shading from the environment
and the low activity of the comet in the vicinity of Philae
-- no detection of particle impacts was made.
Calculations were also performed with the GIADA Per-
formance Simulator (GIPSI) to simulate the expected fluxes
on the DIM instrument during the descent of the lander.
GIADA is an experiment on the Rosetta orbiter devoted
to the measurements of the physical properties of cometary
dust grains in orbit around the nucleus. The grain detec-
tion system (GDS) of GIADA detects the transit of each
single grain entering the instrument by means of a light
curtain. In addition an impact sensor (IS) equipped with
PZT sensors and five microbalances measuring mass de-
position on quartz crystals are included in the experiment
(Colangeli et al. 2007). The Java client software GIPSI is
able to simulate GIADA performance, in particular GIPSI
forecasts how an instrument or a defined surface reacts to
a dust environment along a specific trajectory for defined
time intervals. Inputs to GIPSI are the dust environments
described or evaluated by models (e.g. grain number den-
sity, particle size distribution, and velocity). As inputs to
GIPSI, in addition to the time-dependent 3D model envi-
ronment, we use the spacecraft and comet orbits (the atti-
tude and the position of the spacecraft and the speed along
the orbit for each time step) and instrument parameters
(i.e. field of view, sensitive surface area, subsystems sensi-
tivities). GIPSI simulates the instrument-dust interaction
for each time step (defined as input of the simulation) and
position along the orbit of the spacecraft. The software eval-
uates the vector parameters of the dust model outputs into
the instrument reference frame considering the rotation of
the comet nucleus and the velocity of the spacecraft. The
outputs of the simulation are the number of grains impact-
ing on the surface of the sensor (Della Corte et al. 2014).
In order to simulate the expected fluxes on the DIM
instrument during the descent of the lander we used the
GIPSI simulation software considering as dust environment
the fluxes and the speeds described in Fulle et al. (2010).
The dynamical parameters measured by GIADA during the
same period confirm that the use of Fulle's model is sub-
stantiated (Della Corte et al. 2015; Rotundi et al. 2015).
Article number, page 5 of 17
A&A proofs: manuscript no. AA_postprint
Fulle's dust-tail model (Fulle 1987, 1989) derives an ejec-
tion flux and an ejection velocity for each dust mass (Fulle
1992). The grains are assumed to be spherical. According
to the derived dust ejection velocity, the mass able to es-
cape the nucleus gravity field at a distance of 20 nucleus
radii, assuming a bulk density of 1000 kg m−3, is checked.
Owing to the nucleus asphericity and possibly lower bulk
density (533 kg m−3 as reported by Pätzold et al. (2016)),
the escape velocity is probably significantly lower than the
assumed value of 0.5 m s−1 (Fulle et al. 2010). The method
also requires the value of grain-specific mass to be postu-
lated. The lowest values of the dust mass loss rate is be-
−1 at 3 AU, derived by the assumed
tween 10 and 40 kg s
specific mass and the reported maximum dust mass loss
−1 at 3 AU. The model assigns to all grains
rate 112 kg s
of the same size bin a constant radial velocity equal to the
terminal velocity computed in Fulle et al. (2010). In par-
ticular, we used the upper values for the fluxes reported in
Fulle et al. (2010). We considered an isotropic expansion
of the particle flux, and to calculate the number of parti-
cles emitted, we considered two different densities for the
particles; 100 kg m−3 (fluffy) and 1000 kg m−3 (compact).
Fluffy particles emitted from the comet nucleus were de-
tected in the size range up to a few hundred micrometres
by the Cometary Secondary Ion Mass Analyzer (COSIMA)
on board the Rosetta orbiter (Schulz et al. 2015; Langevin
et al. 2016). From the GDS-only detections of the GIADA
experiment, Fulle et al. (2015) also inferred fluffy particles
of equivalent bulk density of less than 1 kg m−3, which they
associated with the fluffy particles observed by COSIMA.
the
the ESAC repository
spice
(LORL_DL_007_02____P__00220.BSP).
Since
the GIPSI tool is only able to simulate the fluxes over
the Rosetta Orbiter spacecraft, to simulate the fluxes over
the DIM surface we modified the kernel by changing the
reference object of the trajectory and imposing the lander
as reference object. Owing to the lack of a consolidated
spice kernel describing the lander attitude during the
descent we considered an orientation with the +Z side of
the lander parallel to the direction of the force of gravity
(nadir direction) at the given position.
of Philae we
in
trajectory
reported
the
kernel
used
For
The DIM sensor is sensitive to compact particles having
a minimum radius of 0.25 mm and 0.5 mm if the detection
margin is set to 30 dB and 40 dB, respectively. The particles
have a power law size distribution, so it is enough to con-
sider the size bins of particles with radius of 0.21 mm and
0.45 mm in the first approximation. The results of the sim-
ulation for compact particles are shown in Fig. 4. For fluffy
particles the corresponding count rates are only 25% higher
(for comparison of data see Fig. 5). The orientation of Phi-
lae during DIM operational periods was not known at the
time of writing, hence GIPSI fluxes calculated in the nadir
direction are used only as the upper limit. In the GIPSI
tool the configuration of the GDS in GIADA, having a field
of view of 67◦ (corresponding to 1 sr) and a collecting sur-
face of 100 cm2 are considered. This approximation results
in an expected number of counts of maximum 0.005 − 0.5
for compact and for fluffy particles within a total measure-
ment time of 20 minutes. This is in good agreement with
the non-detection of compact particles during descent. Nev-
ertheless, the detection of a fluffy particle with a radius of
1 mm was also an extremely unlikely event. Since the sensor
sides of DIM were pointing off-nadir, a GIPSI calculation
Article number, page 6 of 17
knowing the orientiation of the lander would have provided
even lower values for the number of counts.
6. Conclusions
On the 12 November 2014 lander Philae was deployed from
the Rosetta Orbiter onto the nucleus of comet 67P/C-G.
The DIM experiment of the SESAME instrument package
was switched on several times during descent and on the
surface of the nucleus to detect impacts from submillimetre-
to millimetre-sized particles.
Based on measurements performed with DIM, the upper
limit of the flux of particles in the measurement range of
the instrument was of the order of 10−8 − 10−7m−2s−1sr−1
during descent. The upper limit of the ambient flux of the
submillimetre- and millimetre-sized dust and ice particles at
Abydos was estimated to be 1.6·10−9m−2s−1sr−1 on 13 and
14 November 2014. A correction factor of 1/3 for the field
of view of the sensors was calculated based on an analysis of
the pattern of illumination on Philae. Considering particle
speeds below escape velocity, the upper limit for the volume
density of particles in the measurement range of DIM was
constrained to 10−11 m−3 − 10−12 m−3.
Results of the calculations performed with the GIPSI
tool on the expected particle fluxes during the descent of
Philae are compatible with the non-detection of compact
particles by the DIM instrument.
Acknowledgements. SESAME is an experiment on the Rosetta lan-
der Philae. It consists of three instruments CASSE, DIM, and PP,
which were provided by a consortium comprising DLR, MPS, FMI,
MTA EK, Fraunhofer IZFP, Univ. Cologne, LATMOS, and ES-
TEC. The contribution from MTA EK to the SESAME-DIM ex-
periment was co-funded through the PRODEX contract No. 90010
and by the Government of Hungary through European Space
Agency contracts No. 98001, 98072, 4000106879/12/NL/KML, and
4000107211/12/NL/KML under the plan for European Cooperating
States (PECS). This research was supported by the German Bun-
desministerium für Bildung und Forschung through Deutsches Zen-
trum für Luft- und Raumfahrt e.V. (DLR, grant 50 QP 1302). The
work published in this paper was performed during a visit of A. Hirn
and A. Flandes at MPS. Both guest scientists are grateful to MPS
for financial support during the visit. A. Flandes was also supported
by DGAPA-PAPIIT IA100114 and IA105016. We thank the Rosetta
project at ESA and the Philae project at DLR and CNES for effective
and successful mission operations.
References
Auster, H.-U., Apáthy, I., Berghofer, G., et al. 2015, Science, 349,
Bibring, J.-P., Langevin, Y., Carter, J., et al. 2015, Science, 349,
6247
aab0671
Rev., 128, 205
Sci. Rev., 128, 803
put., 5, 57
583, A13
Bibring, J.-P., Rosenbauer, H., Böhnhardt, H., et al. 2007, Space Sci.
Biele, J., Ulamec, S., Maibaum, M., et al. 2015, Science, 349, 6247
Colangeli, L., Lopez-Moreno, J. J., Palumbo, P., et al. 2007, Space
Della Corte, V., Ivanovski, S., Lucarelli, F., et al. 2014, Astron. Com-
Della Corte, V., Rotundi, A., Fulle, M., et al. 2015, Astron. Astrophy.,
Fischer, H.-H. 2012, SESAME Flight Software FM-3 -- Telecommands
and Telemetry Formats, Rosetta technical report RO-LSE-UG-
3404, version 5.5a
Fischer, H.-H. 2014, Considerations on DIM Detection Efficiency
(Single Impacts) and the Current Situation of DIM Calibration,
SESAME document, unpublished
Flandes, A., Krüger, H., Loose, A., Albin, T., & Arnold, W. 2014,
Planet. Space Sci., 99, 128
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
122
Flandes, A., Krüger, H., Loose, A., et al. 2013, Planet. Space Sci., 84,
Fulle, M. 1987, Astron. Astrophys., 171, 327
Fulle, M. 1989, Astron. Astrophys., 217, 283
Fulle, M. 1992, Astron. Astrophys., 265, 817
Fulle, M., Colangeli, L., Agarwal, J., et al. 2010, Astron. Astrophys.,
522, A63
Lett., 802, L12
92, 156
Astrophy., 583,
271, 76
Fulle, M., Della Corte, V., Rotundi, A., et al. 2015, Astrophys. J.
Glassmeier, K.-H., Böhnhardt, H., Koschny, D., Kührt, E., & Richter,
I. 2007, Space Sci. Rev., 128, 1
Hertz, H. 1882, Journal über die reine und angewandte Mathematik,
Johnson, K. L. 2003, Contact Mechanics (Cambridge University Press,
ISBN 0-521-34796-3), 351 -- 360
Krüger, H., Seidensticker, K. J., Fischer, H.-H., et al. 2015, Astron.
Langevin, Y. and, H. M., Ligier, N., Merouane, S., et al. 2016, Icarus,
Pätzold, M., Andert, T., Hahn, M., et al. 2016, Nature, 530, 63
Péter, A. 2002, Performance Report of DIM, Rosetta technical report
RO-LSE-TP-3445, Issue 1
Péter, A. & Lucás, V. 2001, DIM Software Description, SESAME
document RO-LSE-SP-3440, version 4
Rotundi, A., Sierks, H., Della Corte, V., et al. 2015, Science, 347,
Schulz, R., Hilchenbach, M., Langevin, Y., et al. 2015, Nature, 518,
Seidensticker, K. J., Möhlmann, D., Apáthy, I., et al. 2007, Space Sci.
Spohn, T., Knollenberg, J., Ball, A. J., et al. 2015, Science, 349,
aaa3905
216
Rev., 128, 301
aab0464
Thomas, G. R. & Willis, D. M. 1972, J. Phy. E, 5, 260
Article number, page 7 of 17
List of Figures
A&A proofs: manuscript no. AA_postprint
1
5
DIM on the top of Philae. The sensor is visible in the corner (Krüger et al. 2015). Credits: ESA/ATG
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
medialab.
2 Maximum angles of incidence at which the impacting particles are not shaded by the frame of the sensor.
Diagram of the basic geometry used in the simulation explained in Sect. 4.3. The lander Philae is sur-
3
rounded by a barrier curved as a sinusoid. The front (bottom of the image) and the space above the lander
are open. The illumination of the Sun mainly comes from the front because the Sun moves on a plane that
forms < 22◦ with the floor of this box. The ≈ 90◦ angle indicates the width of the field of view of the
lander towards the Sun.
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Count rates of compact particles in the size bins with radius 0.21 mm and 0.45 mm as calculated with
GIPSI for a collecting surface area of 100 cm2 and a field of view of 67◦ (GDS configuration).
. . . . . . .
Count rates of compact and fluffy particles in the size bin of particles with radius 0.45 mm as calculated
with GIPSI for a collecting surface area of 100 cm2 and a field of view of 67◦ (GDS configuration).
. . . .
4
9
10
11
12
13
Article number, page 8 of 17
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
Fig. 1. DIM on the top of Philae. The sensor is visible in the corner (Krüger et al. 2015). Credits: ESA/ATG medialab.
Article number, page 9 of 17
XYZSD2DIMSolar Array!Panel 1A&A proofs: manuscript no. AA_postprint
Fig. 2. Maximum angles of incidence at which the impacting particles are not shaded by the frame of the sensor.
Article number, page 10 of 17
82ºincoming particleYZXd=2.3 mmL=50 mmW=16 mm87ºincoming particleYXZθφA. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
Fig. 3. Diagram of the basic geometry used in the simulation explained in Sect. 4.3. The lander Philae is surrounded by a barrier
curved as a sinusoid. The front (bottom of the image) and the space above the lander are open. The illumination of the Sun mainly
comes from the front because the Sun moves on a plane that forms < 22◦ with the floor of this box. The ≈ 90◦ angle indicates the
width of the field of view of the lander towards the Sun.
Article number, page 11 of 17
Sinusoid-shaped barrier~ 90°to open sideA&A proofs: manuscript no. AA_postprint
Fig. 4. Count rates of compact particles in the size bins with radius 0.21 mm and 0.45 mm as calculated with GIPSI for a collecting
surface area of 100 cm2 and a field of view of 67◦ (GDS configuration).
Article number, page 12 of 17
00:0004:0008:0012:0016:0020:0024:000123 bin: 4.54(cid:215)10-4 bin: 2.11(cid:215)10-4count rate (particles h-1)time (hh:mm)12-11-2014DIM detection at 14:43:47 UTCA. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
Fig. 5. Count rates of compact and fluffy particles in the size bin of particles with radius 0.45 mm as calculated with GIPSI for
a collecting surface area of 100 cm2 and a field of view of 67◦ (GDS configuration).
Article number, page 13 of 17
00:0004:0008:0012:0016:0020:0024:000,00,10,20,30,40,50,60,70,812-11-2014compact; bin: 4.54(cid:215)10-4fluffy; bin: 4.50(cid:215)10-4count rate (particles h-1)time (hh:mm)DIM detection at 14:43:47 UTCA&A proofs: manuscript no. AA_postprint
List of Tables
1
2
3
Geometric factors G calculated for the three sensor sides of DIM . . . . . . . . . . . . . . . . . . . . . . .
Upper limit of the ambient particle flux during descent (12 Nov 2014)
. . . . . . . . . . . . . . . . . . . .
Particle flux calculated for the three sensor sides of DIM after landing at Abydos . . . . . . . . . . . . . .
15
16
17
Article number, page 14 of 17
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
Table 1. Geometric factors G calculated for the three sensor sides of DIM
Model for the sensor side
3 stand-alone PZT segments (3G0)
3 PZT segments with frame (3Gf)
Lander configuration
(relative to 3G0)
(detection ratio)
X
76.3
63.7
34.0
(-56%)
(0.445)
Geometric factor (cm2sr)
Y
76.3
63.7
47.9
(-37%)
(0.628)
Z
76.3
63.7
51.0
(-33%)
(0.668)
Article number, page 15 of 17
Table 2. Upper limit of the ambient particle flux during descent (12 Nov 2014)
A&A proofs: manuscript no. AA_postprint
Dist. DIM Meas. Eff. meas. Margin Meas.
range
(mm)
(8)
side
Meas.
numb.
(1)
30
31
32
33
34
35
36
37
38
39
40
41
Start
time
(UTC)
(2)
08:38:32
08:42:23
08:46:13
08:50:03
08:53:52
08:57:42
09:59:04
10:02:54
10:06:44
14:40:04
14:42:14
14:44:24
(km)
(3)
22.2
22.1
22.0
21.8
21.6
21.4
18.6
18.4
18.3
5.1
5.0
4.9
time
(s)
(5)
200
200
200
200
200
200
200
200
200
100
100
100
(4)
X
Y
Z
X
Y
Z
X
Y
Z
X
Y
Z
time
(s)
(6)
0
0
198
200
200
200
0
198
0
96
98
0
(dB)
(7)
40
40
40
50
50
50
40
40
40
40
40
40
0.5 - 6.5
0.5 - 6.5
0.5 - 6.5
0.9 - 6.5
0.9 - 6.5
0.9 - 6.5
0.5 - 6.5
0.5 - 6.5
0.5 - 6.5
0.5 - 6.5
0.5 - 6.5
0.5 - 6.5
Det.
events
(9)
0
0
0
0
0
0
0
0
0
0
1
0
Nmax.
(s−1)
(10)
--
--
--
--
--
1.5 · 10−2
1.5 · 10−2
1.5 · 10−2
1.5 · 10−2
1.5 · 10−2
3.1 · 10−2
4.8 · 10−2
Φmax.
(m−2s−1sr−1)
(11)
--
--
--
--
--
3 · 10−8
4 · 10−8
3 · 10−8
3 · 10−8
3 · 10−8
9 · 10−8
1 · 10−7
Article number, page 16 of 17
A. Hirn et al.: Estimation of particle flux based on SESAME-DIM measurements
Table 3. Particle flux calculated for the three sensor sides of DIM after landing at Abydos
Sensor side Meas. time
+X
+Y
+Z
(s)
5579
5579
5579
Nmax.
(s−1)
5.4 · 10−4
5.4 · 10−4
5.4 · 10−4
(m−2s−1sr−1)
Φmax.
1.6 · 10−9
1.1 · 10−9
1.1 · 10−9
Notes. The measurement times of the individual measurement blocks during FSS were reported by Krüger et al. (2015).
Article number, page 17 of 17
|
1109.1614 | 1 | 1109 | 2011-09-08T02:37:01 | Are there rings around Pluto? | [
"astro-ph.EP"
] | Considering effects of tidal plus centrifugal stress acting on icy-rocks and the tensile strength thereof, icy-rocks being in the density range (1-2.4) g cm-3 which had come into existence as collisional ejecta (debris) in the vicinity of Pluto at the time when Pluto-Charon system came into being as a result of a giant impact of a Kuiper Belt Object on the primordial Pluto, it is shown, here, that these rocks going around Pluto in its vicinity are under slow disruption generating a stable ring structure consisting of icy-rocks of diameters in the range (20-90) km, together with fine dust and particles disrupted off the rocks, and spread all over the regions in their respective Roche Zones, various Roche radii being in ~1/2 three-body mean motion resonance. Calculations of gravitational spheres of influence of Pluto which turns out to be 4.2 x 106 km for prograde orbits and 8.5 x 106 km for retrograde orbits together with the existence of Kuiper Belt in the vicinity of Pluto assure that there may exist a few rocks (satellites)/dust rings/sheets so far undiscovered moving in prograde orbits around the planet and few others which are distant ones and move around Pluto in the region between 4.2x106 km and 8.5x106 km in retrograde orbits. | astro-ph.EP | astro-ph | IJFPS, Vol.1, No.1, pp. 6-10, June, 2011
J.J.Rawal
Are there rings around Pluto?
J. J. Rawal 1 , Bijan Nikouravan 2, 3
1The Indian Planetary Society (IPS) B-201, Lokmanya Tilak Road, Borivali (W), Mumbai – 400092, India
Email: [email protected]
2 Department of Physic, Islamic Azad University (IAU) -Varamin - Pishva Branch, Iran
3University of Malaya, 50603 Kuala Lumpur, Malaysia
Email: [email protected]
(Received May 2011; Published June 2011)
ABSTRACT
Considering effects of tidal plus centrifugal stress acting on icy-rocks and the tensile strength thereof, icy-rocks being in the
density range (1–2.4) g cm-3 which had come into existence as collisional ejecta (debris) in the vicinity of Pluto at the time when
Pluto-Charon system came into being as a result of a giant impact of a Kuiper Belt Object on the primordial Pluto, it is shown,
here, that these rocks going around Pluto in its vicinity are under slow disruption generating a stable ring structure consisting of
icy-rocks of diameters in the range (20–90) km, together with fine dust and particles disrupted off the rocks, and spread all over
the regions in their respective Roche Zones, various Roche radii being in ~1/2 three-body mean motion resonance. Calculations of
gravitational spheres of influence of Pluto which turns out to be 4.2 x 10 6 km for prograde orbits and 8.5 x 106 km for retrograde
orbits together with the existence of Kuiper Belt in the vicinity of Pluto assure that there may exist a few rocks (satellites)/dust
rings/sheets so far undiscovered moving in prograde orbits around the planet and few others which are distant ones and move
around Pluto in the region between 4.2x106 km and 8.5x106 km in retrograde orbits.
Key words: Giant impact-Pluto-Charon System- Roche-limit- resonance- rings- new satellites- Kuiper-Belt.
INTRODUCTION
Now that Pluto is no longer considered a planet and has been
known to have three satellites, namely, Charon (Christy and
Harrington, 1978, 1980; Harrington and Harrington, 1980;
Harrington and Christy 1980a,b), Nix and Hydra (Weaver et al.,
2006), a question has arisen as to whether Pluto has a ring
structure and satellites so far undiscovered going around it.
That NASA’s robotic space-probe “New Horizons” has been on
its way to Pluto, several authors in trying to answer the
question, have analyzed the problem and would go on
analyzing it untill the space- probe will reach Pluto in the year
2015. Some planetary scientists (Stern 1988, 1995, 2002; Stern
et al. 1991, 1994, 1997a, b, 2003, 2006a, b; Steffl and Stern
2007; Steffl et al., 2006; Tholen and Buie 1988, 1990, 1997a,
b) have come out with the prediction that Pluto may have a
time-variable ring / dust sheets / partial rings around it which
6
are far away from the planet. These rings were/are formed due
to collisions of Kuiper Belt Objects with Nix and Hydra. Here,
in this paper, we attempt to show that Pluto may have a stable
ring structure and a
few more satellites (rocks), yet
undiscovered going around it.
ROCHE LIMIT
The concept of Roche limit is well- known in the literature
(Jeans 1928, 1960; Jeffery’s, 1947) . For a rigid body of density
𝜌𝑠 revolving around a primary having radius R and density 𝜌 ,
the Roche limit around the primary with respect to such a
secondary is given by
1/3
𝜌
𝜌𝑠
(1)
𝑅
𝑅𝑅𝑜𝑐 ℎ 𝑒 ≈ 1.44
IJFPS, Vol.1, No.1, pp. 6-10, June, 2011
In what follows, we shall use system parameters as given in
Table 1.
Table 1. Parameters for Pluto–Charon System that are used here.
* There is some uncertainty in the values of densities of Pluto and Charon.
VARIOUS HYPOTHESES FOR FORMATION OF
PLUTO-CHARON SYSTEM
After the discovery of Pluto’s satellite Charon, several
authors, (Mignard, 1981; Lin, 1981; McKinnon, 1984, 1988,
1989; Mckinnon and Muller 1988; Canup 2005) , in trying to
understand the formation of Pluto-Charon system, put forth
several hypotheses such as fission and a giant impact origin.
According to (Mignard 1981), there are numerous observed and
theoretical facts which favour an origin of Charon by fission.
Several authors have investigated the possibility for Pluto to be
an ejected satellite of Neptune (Lyttleton 1936, 1953, 1961;
Hoyle, 1975; Harrington and Van Flandern, 1979; and
Dormand and Woolfson, 1980). After a so violent event the
new angular speed of Pluto can give birth to a rotational
instability which leads to the break-up of the primordial Pluto.
If it was so, the radius of Pluto over that of Charon is to be
close to 1.9. This result is in keeping with the new philosophy
which is currently emerging in the field of planetology. The
solar system is definitely more diversified than it was thought.
Then, there is no reason to believe in a unique process to
originate satellites, and mechanisms as varied as accretion,
capture, fission and giant impact origin have likely been
efficient throughout the Solar System.
Lin (1981) proposed that Pluto-Charon System might have
been formed by binary fission of a rapidly rotating body. If, as
a result of fission, Charon was formed just inside its own tidal
radius, the mass ratio of Charon to Pluto must not exceed 0.25.
Otherwise the resultant spin angular momentum of Pluto would
cause it to break up again. The mass ratios must be greater than
0.05 in order for Charon to form outside the unstable co -
rotation radius and subsequently evolve to the present stable co-
rotation radius rather than be driven back to Pluto. The
observational value for mass ratios is 0.1, which is consistent
with the limit set by his own hypothesis (This was the case in
1981 when the correct mass ratio of Charon to Pluto and other
Pluto-Charon System parameters were exactly unknown, even
today also some system parameters are not known exactly).
Furthermore, it is very close to the critical value for mass ratio
in which the initial object has enough angular momentum to
become secularly unstable. Lin (1981) felt
that further
observations on the mass, size, binary separation and the
density of Pluto and Charon would provide new insight into the
7
J.J.Rawal
process of binary fission and planetary formation (also see
Foust et al., 1997).
Precise determination of diameter of Pluto and Charon along
with the total mass of the system provide a powerful basis for
constraining the origin of this enigmatic planetary pair. The
work of Mckinnon (1989) focuses on the angular momentum
budget of Pluto-Charon, taking as the point of departure from
earlier work of Lin (1981). Because of the large angular
momentum density of the system, he argued for an impact or
collisional origin. He then addressed aspects of the required
impact process and compared them with somewhat similar
hypotheses for the Moon’s origin. It has been recognized for
some time that Pluto-Charon’s J value is high, but it was not
known to be very high as it has been found to be now. Lin
(1981) and Mignard (1981), therefore, advocated fission of a
single original object. McKinnon (1984), Burns (1986) and
Peale (1986) suggested a large-body impact. For comparison,
the Earth-Moon system has a J of 0.115, and even this has been
judged great enough for impact over spinning to be seriously
considered [Durisen and Gingard, (1986)]. The logical cause of
Pluto-Charon’s large J is a large-body impact. An impact origin
is physically plausible as it is for suspected binary asteroids
(Weidenschilling et al., 1989, Weidenschilling 2002; see also
Mckinnon and Muller, 1988).
Canup
to
(2005) used hydro dynamical simulations
demonstrate that the formation of Pluto-Charon by means of a
large collision is quite plausible. He also observed that such an
impact probably produced an intact Charon, although it is
possible that a disc of material orbited Pluto from which
Charon later accumulated. These findings suggested that
collisions between 1000-kilometer-class objects occurred in the
early inner Kuiper Belt.
ROLE OF A GIANT IMPACT ORIGIN HYPOTHESIS,
ROCHE LIMIT AND A THREE-BODY MEAN MOTION
RESONANCE IN THE FORMATION OF A STABLE
RING STRUCTURE AROUND PLUTO
According to a Giant Impact Origin Hypothesis, Pluto -
Charon System came into being as a result of a collision of a
big Kuiper Belt Object (1000 km size) with primordial Pluto. If
this is so, it is natural to expect collisional ejecta (debris,
fragments) spread all around in the Pluto -Charon System and
revolve around Pluto-the biggest of all remnants of the
catastrophic collisional event. This shows that the region in the
Pluto-Charon System may not be clean but would be full of
small or big collisional ejecta (debris, fragments in the form of
rocks) and also the facts that the two newly discovered
satellites Nix and Hydra of Pluto (Weaver et al.2006) are in
proximity to Pluto and Charon, they are on near-circular orbits
in the same plane as Pluto’s large satellite Charon, along with
their apparent locations in or near high order mean motion
resonances, all probably result from their being constructed
from collisional ejecta that originated from the Pluto-Charon
System formation event (Stern et al., (2006a, b). Stern et al.,
(2006a, b) also argue that dust rings of variable optical lengths
IJFPS, Vol.1, No.1, pp. 6-10, June, 2011
J.J.Rawal
form sporadically in the Pluto System far away from Pluto due
to collisions of Kuiper Belt objects with Nix and Hydra and that
rich satellite systems may be found perhaps frequently-around
other large Kuiper belt objects. Let us, therefore, consider rocks
(satellites) having densities in the range (1–2.4) g cm-3 (Table
2) lying in the neighborhood of Pluto. We now calculate the
Roche limits, 𝑅𝑖 , with respect to rocks of densities in the range
(1–2.4) g cm-3. These are shown in (Table 2). Taking tensile
strength, s, for an ice-rock to be 3×106 dyne cm-2 (Jeffreys,
1947) in the formula,
1/2
19𝑎 3 𝑠
32 𝐺𝑀 𝜌𝑠
𝑟 =
(2)
Where 𝑟 is the reduced radius of a satellite upto which the
satellite has been ruptured; a, the radial distance at which the
satellite (rock) is ruptured to its maximum; G, the Universal
gravitational constant; M, the mass of Pluto; 𝜌𝑠 , the density of
satellite and s, is the tensile strength. We calculate the reduced
radius up to which each rock has been ruptured at various
Roche radii Ri, rock density 𝜌𝑠𝑖 being in the range (1–2.4) g cm-
3. This is shown in Table 2.
Table 2. Location of Stable Ring Structure around Pluto. Table showing Roche radial distances, minimum radii that satellites can retai n at corresponding Ri and
when they graze the planet; revolution periods, Ti, corresponding to Ri. All these parameters corresponding to densities, Si, in the range (1―2.4) g cm-3
Densities, si of
rocks(secondary) in
(g cm-3)
Corresponding
Roche limit Ri (km)
Minimum radius r (km) that
a satellite can retain at
corresponding Ri
Minimum radius r (km)
That a satellite can retain
when it grazes the planet.
Ti [revolution period
in days at corresponding
Roche radial distance (Ri)]
s 8 = 1
s 7 = 1.2
s 6 = 1.4
s 5 = 1.6
s 4 = 1.8
s 3 = 2.0
s 2 = 2.2
s 1 = 2.4
R8 = 2149.5
R7 = 2022.1
R6 = 1920.9
R5 = 1837.2
R4 = 1767.7
R3 = 1706.0
R2 = 1653.5
R1 = 1604.1
r8 = 43.55
r7 = 36.28
r6 = 31.28
r5 = 27.21
r4 = 24.21
r3 = 21.72
r2 = 19.81
r1 = 18.12
r8 = 16
r7 = 14.6
r6 = 13.5
r5 = 12.6
r4 = 11.9
r3 = 11.3
r2 = 11.0
r1 = 10.3
T8 = 0.2369
T7 = 0.2162
T6 = 0.2001
T5 = 0.1872
T4 = 0.1766
T3 = 0.1675
T2 = 0.1598
T1 = 0.1528
From Table 2, it is clear that the reduced radii up to which the
rocks have been ruptured are in the range 15 to 45 km with
respect to Roche radii, and are in the range 10 to 16 km at the
radial distance grazing the planet. All these rocks lie within the
radial distance ~2,500 km around Pluto. This, we believe, form
a ring around Pluto. As tidal disruption goes on in this region,
the region is full of fine dust, particles and small or big pebbles
and rocks, forming a stable ring structure around Pluto.
2𝜋
Resonance theory states that if 𝑛1 , 𝑛2 , 𝑛3 , (𝑛𝑖 =
, 𝑛1 . > 𝑛2 >
𝑇
𝑛3 ) ,are mean motions of three secondaries in circular orbits,
then condition for frequent occurrence of mirror configuration
(Dermott 1968 a, b, 1973; Greenberg, 1973; Gold reich, 1965 a,
b; Ovenden et al., 1974; Roy and Ovenden, 1955; Rawal 1981,
1989; Goldreich and Soter, 1966; and references given therein)
is given by the Equation
𝛼𝑛1 − 𝛼 + 𝛽 𝑛2 + 𝛽𝑛3 = 0
(3)
Where α, β are small and mutually prime positive integers. It
follows from Equation (3) that in a reference frame rotating
with the mean motion of any one of the three secondaries, the
′ of other two are commensurate, and
relative mean motions 𝑛𝑖
that in a frame I (that of the innermost secondary), we have
′
𝑛 2
′ =
𝑛 3
𝑛 2−𝑛 1
𝑛 3−𝑛 1
=
𝛽
𝛽 +𝛼
(4)
In terms of revolution periods, Equation (4) is written as
′
𝑛 2
′ =
𝑛 3
𝑇3 𝑇2−𝑇1
𝑇2 𝑇3−𝑇1
=
𝛽
𝛽 +𝛼
(5)
8
In order to know whether a triad of successive secondaries at
various Roche radial distances Ri (given in Table 2) is in stable
three body mean motion resonance, we calculate revolution
periods Ti (shown in Table 2) corresponding to various Ri to
find corresponding mean motions. From Equation (6) we find
that a triad of successive secondaries at various Ri has relative
mean motion ratio ~1/2 throughout, that is, they follow three -
body mean motion resonance relation given by
𝑛1 − 2𝑛2 + 𝑛3 = 0
(6)
This shows that various rocks (satellites) at or near these
resonance orbits are in reasonably stable resonant orbits, and if
they existed there, then they still exist there. In other words,
rocks and particles along with fine dust which got disrupted off
the rocks form a stable ring structure in the vicinity of Pluto
which is within the distance ~2500 km from the centre of the
planet.
GRAVITATIONAL SPHERE OF
INFLUENCE OF
PLUTO FOR PROGRADE AND RETROGRADE
ORBITS
AND
THEIR
SIGNIFICANCE
FOR
EXISTENCE OF UNKNOWN SATELLITES IN THE
SYSTEM.
In order to know the existence of distant unknown satellites
in the Pluto- Charon System, we would like to know how far
the gravitational influence of Pluto is, that is, how large the
gravitational sphere of influence of Pluto is. We, therefore,
calculate here, the boundaries to the gravitational sphere of
IJFPS, Vol.1, No.1, pp. 6-10, June, 2011
J.J.Rawal
influence of Pluto for prograde and retrograde orbits using
King- Innanen formula. Innanen (1979) modified the King
(1962) formula for putting in the equation for acceleration in a
revolving co-ordinate frame with an additional Coriolis term of
magnitude , 2Ω𝑣𝑟 where 𝑣𝑟 is the velocity of the secondary
relative to the primary. Here in our case, we consider Moon-
Pluto-Sun System, and therefore, the primary is the Sun and the
secondary is Pluto. The familiar right hand rule immediately
shows that the Coriolis term is always directed radially between
the secondary (Pluto) and the primary (Sun). It counteracts the
primary’s gravity for the direct motion of the moon, but
effectively supplements the primary’s gravity for retrograde
motion. For the limiting direct and retrograde radii, r d and rr
respectively, of a moon around a planet (here Pluto) in the most
general case where
the planet (here Pluto)’s orbit has
eccentricity e and the pericentric distance 𝑅 = 𝑅𝑝 = 𝑎(1 − 𝑒), a,
being the mean distance of the planet (Pluto), we have
𝑟𝑟
𝑟 𝑑
Where
= 𝑓(𝑒) 2/3
(7)
1
𝑓 (𝑒 ) 2
𝑚
𝑀
1
3 𝑅𝑝
𝑟𝑑 =
(8)
5+𝑒+2(4+𝑒 )1/2
3+𝑒
(9)
𝑓 𝑒 =
Here, 𝑚, is the mass of the planet (here Pluto), M, is the mass
of the Sun. Therefore, calculation for the gravitational sphere of
influence of Pluto for prograde orbits turns out to be 4.2 × 106
km and for retrograde orbits, it is 8.5 × 106 km. The distances
of Charon, Nix, Hydra are 19300, 48675, 64780 km
respectively. These satellites are very much inside the boundary
of the gravitational sphere of influence for prograde satellites.
The boundary of the gravitational sphere of influence for
retrograde satellites is far far away.
At the time of a giant impact, collisional ejecta (debris) were
likely to be thrown far away. It is, therefore, likely that there
may be few prograde satellites (rocks) revolving around Pluto,
between Pluto and Charon and beyond Hydra within the
distance 4.2 × 106 km, and a few retrograde satellites (rocks)
revolving around the planet between the distance 4.2 × 106 km
and 8.5 × 106 km. Moreover, in the vicinity of Pluto, there is
Kuiper belt. It is likely that some Kuiper belt Objects might
have been captured by Pluto making them its satellites and / or
REFERENCES
Burns, J. A. 1986, in Satellites, eds. J. A. Burns and M. S.
Matthews (Tucson: University of Arizona Press) p. 117.
Canup, R. M. 2005, Science 307, 546.
Christy, J. W., and Harrington R. S. 1978, Astr. J. 83, 1005.
Christy, J. W., and Harrington R. S. 1980, Icarus 44, 38.
Dermott, S. F. 1968a, Mon. Not. R. Astr. Soc. 141, 349.
Dermott, S. F. 1968b, Mon. Not. R. Astr. Soc. 141, 363.
9
injected into the gravitational sphere of influence of Pluto
thereby becoming its satellites. They may be prograde moving
if they had entered into the gravitational sphere of influence of
Pluto for prograde satellites, that is, within the distance
4.2 × 106 km, and may be retrograde moving if they are
outside it, but inside the gravitational sphere of influence of
Pluto for retrograde motion, that is, distance between 4.2 × 106
km and 8.5 × 106 km. This analysis shows that Pluto may have
a few, yet undiscovered satellites going around it, a few of them
even moving in retrograde direction.
CONCLUSION
Here, it is shown that a stable ring system consisting of small
rocks having densities in the range (1–2.4) g cm-3 and diameters
in the range (20–90) km along with fine dust and particles
disrupted off these rocks, may exist around Pluto within the
distance ~2500 km from the centre of the planet. There may
also exist a few satellites (rocks) other than already known. If
these satellites (rocks) orbit the planet within the distance
4.2 × 106 km, then they are protrude moving, and if they orbit
the planet in the region between 4.2 × 106 and 8.5 × 106 km,
then they are retrograde moving. These satellites may be
collisional ejecta (debris) which came into existence as a result
of catastrophic collision event which formed Pluto-Charon
System or rocks captured by Pluto from Kuiper belt and made
them its own satellites or Kuiper belt Objects injected into
Pluto’s gravitational sphere of influence by some gravitational
perturbations due to Neptune or Oort’s cometary cloud as a
whole becoming its satellites
ACKNOWLEDGEMENTS
Thanks are due
to Professors K. Sivaprasad and S.
Ramadurai (both retired from Tata Institute of Fundamental
Research, Mumbai)
for helpful discussions and useful
suggestions and Mr. Hemal S. Shah, Mr. D. K. Sahani and Mr.
Darshan K. Vishwakarma for their technical assistance.
Thanks are also due to anonymous referees for their
constructive criticisms and helpful suggestions that have greatly
improved the content of this paper.
Dermott, S. F. 1973, Nature 244, 18.
Durisen, R. H., and Gingard R. A. 1986, in Origin of the
Moon, eds. W. K. Hartmann, R. J. Phillips and G. J. Taylor
(Houston: Lunar and Planetary Institute), p.487.
Dormand, J. R. and Woolfson M. M. 1980, Mon. Not. R.
Astr. Soc. 193, 171.
Foust, Jeffrey A., Elliot, J. L., Olkin, Catherine B.,
McDonald, Stephen W., Dunham, Edward W., Stone,
Remington P. S., McDonald John S., and Stone, Ronald C.
1997, Icarus 126, 362.
Goldreich, P. 1965a, Mon. Not. R. Astr. Soc., 130, 159.
IJFPS, Vol.1, No.1, pp. 6-10, June, 2011
J.J.Rawal
Goldreich, P. 1965b, Astr. J. 70, 5.
Goldreich, P. & Soter, S. 1966, Icarus 5,375.
Greenberg, R. J. 1973, Astr. J. 78, 338.
Harrington, R. S. & Van Flandern, T. C. 1979, Icarus 39,
131.
Harrington,R. S. and Harrington, B. J. 1980, Sky and
Telescope 59, 452.
Harrington, R. S. & Christy, J. W. 1980a, Astr. J., 85, 168.
Harrington, R. S. & Christy, J. W. 1980b, IAU circ.3515.
Hoyle, F. 1975, Astronomy and Cosmology, W. H. Freeman
and Co. San Franscisco, pp. 441-451.
Innanen, K. A 1979, A. J. 84, 960.
Jeans, J. N. 1928, Astronomy and Cosmogony, Cambridge
University Press.
Jeans, J. 1960, Universe Around Us, 4 th edition, Cambridge
University Press.
Jeffreys, H. 1947, Mon. Not. R. Astr. Soc. 107, 260.
King, I. R. 1962, Astr. J. 67, 471.
Laplace, P. S. de 1805, Mecanique Celeste, Vol. 4, Courier,
Paris, Trans. N. Bowditch rpt. 1966, New York, Chelsea.
Lin, D. N. C. 1981, Mon. Not. R. Astr. Soc. 197, 1081.
Lyttleton, R. A. 1936, Mon. Not. R. Astr. Soc. 97, 108.
Lyttleton, R. A. 1953, The Stability of Rotating Liquid
masses, Cambridge University Press.
Lyttleton, R. A. 1961, Man’s View of the Universe, Micheal
Joseph, London.
McKinnon, W. B. 1984, Nature, 311, 355.
McKinnon, W. B. and Muller S. 1988, Nature 335, 240.
McKinnon, W. B. 1989, Ap. J., 344, L 41-L 44.
Mignard, F. 1981, Astrn. Astrophys. 96, L 1-L 2.
Ovenden, M. W. Feagen, T., Grab, O. 1974, Cel. Mech. 8,
455.
Peale, S. J. 1986, in satellites, eds. J. A., Burns & M. S.
Matthews (Tucson; Univ. Arizon Press), p. 159.
Rawal, J. J. 1981, Moon and Planets, 24, 407.
Rawal, J. J. 1989, J. Astrophys. Astr. 10, 257.
Roy, A. E. and Ovenden, M. W. 1955, Mon. Not. R. Astr.
Soc. 115, 295.
Steffl, A. J., Mutchler, M. J., Weaver H. A., Stern, S. A.,
Durda, D. D., Terrell, D., Merline W. J., Young, L. A., Young
E. F., Buie, M. W., and Spencer J. R. 2006, Astr. J. 132, 614.
Steffl, A. J. and Stern, S. A. 2007, Astr. J. 133, 1485.
Stern, S. A. 1988, Icarus 73, 269.
Stern, S. A. Parker, J. Wm., Fesen, R. A., Barker, E. S. &
Trafton, L. M. 1991, Icarus 94, 246.
Stern, S. A., Parker, J. W., Duncan, M. J., Snowdall, J. C., Jr.,
& Levison, H. F. 1994, Icarus, 108, 234.
Stern, S. A. 1995, Astr. J. 110,856.
Stern, S. A., Buie, M. W., & Trafton, L. M. 1997a, AJ, 113,
827.
Stern, S. A., McKinnon, W. B, & Lunine, In Pluto And
Charon, eds., S. A. Stern and D. J. Tholen, 605 U. Ariz. Press
1997b.
Stern, S. A. 2002, Astr. J. 124, 2300.
Stern, S. Alan, Bottke, William F. And Lerison, Harold F.
2003, AJ 125, 902.
Stern, S. A., Weaver, H. A., Steffl, A. J., Mutchler, M.,
Merline, W. J., Buie, M. W., Young E. F., Young, L. A., &
Spencer, J. R., 2006a, Nature 439, 946.
Stern, S. A., Mutchler, M. J., Weaver, H. A., & Steffl, A. J.
2006b, Arxiv Astrophysics e-prints.
Tholen, D. J., and Buie, M. W., 1988, A.J. 96, 1977.
Tholen, D. J., and Buie, M. W., 1990, Bull. Am. Astron. Soc.
22, 1129.
Tholen, D. & Buie, M. 1997a, Pluto and Charon, eds. S. A.
Stern & D. Tholen (Tucson, U. Az. Press), p.193.
Tholen, D. J., and Buie M. W. 1997b, Icarus 125, 245.
Weaver, H. A., Stern, S.A., Mutchler, M. J.,Steffl, A. J., Buie
M.W., Merline, W. J., Spencer, J. R., Young E. F., & Young, L.
A. 2006, Nature, 439,943.
Weidenschilling, S. J., Paolicchi, P., and Zappala, V. 1989, in
Asteroids II, eds. R. P. Binzel, T., Gehrels, and M. S. Matthews
(Tucson: University of Arizona Press).
Weidenschilling, S. J. 2002, Icarus, 108, 219.
10
|
1104.3804 | 1 | 1104 | 2011-04-19T17:10:59 | Precision radial velocities with CSHELL | [
"astro-ph.EP",
"astro-ph.SR"
] | Radial velocity identification of extrasolar planets has historically been dominated by optical surveys. Interest in expanding exoplanet searches to M dwarfs and young stars, however, has motivated a push to improve the precision of near infrared radial velocity techniques. We present our methodology for achieving 58 m/s precision in the K band on the M0 dwarf GJ 281 using the CSHELL spectrograph at the 3-meter NASA IRTF. We also demonstrate our ability to recover the known 4 Mjup exoplanet Gl 86 b and discuss the implications for success in detecting planets around 1-3 Myr old T Tauri stars. | astro-ph.EP | astro-ph |
Precision radial velocities with CSHELL
Christopher J. Crockett 1,2,3,
Naved I. Mahmud 3,4,
L. Prato 1,3,
Christopher M. Johns-Krull 3,4,
Daniel T. Jaffe 5,
Charles A. Beichman 6,7
Received
;
accepted
1Lowell Observatory, 1400 W Mars Hill Road, Flagstaff, AZ 86001;
[email protected], [email protected]
2Department of Physics and Astronomy, University of California Los Angeles,
430 Portola Plaza, Box 951547, Los Angeles, CA 90095-1547
3Visiting Astronomer at the Infrared Telescope Facility, which is operated by the Univer-
sity of Hawaii under Cooperative Agreement no. NCC 5-538 with the National Aeronautics
and Space Administration, Science Mission Directorate, Planetary Astronomy Program.
4Department of Physics and Astronomy, Rice University, MS-108, 6100 Main Street,
Houston, TX 77005; [email protected], [email protected]
5Department of Astronomy, University of Texas, R.L. Moore Hall, Austin, TX 78712;
[email protected]
6Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive,
Pasadena, CA 91109
7NASA Exoplanet Science Institute (NExScI), California Institute of Technology, 770 S.
Wilson Ave, Pasadena, CA 91125
-- 2 --
ABSTRACT
Radial velocity identification of extrasolar planets has historically been dom-
inated by optical surveys. Interest in expanding exoplanet searches to M dwarfs
and young stars, however, has motivated a push to improve the precision of near
infrared radial velocity techniques. We present our methodology for achieving
58 m s−1 precision in the K band on the M0 dwarf GJ 281 using the CSHELL
spectrograph at the 3-meter NASA IRTF. We also demonstrate our ability to
recover the known 4 MJU P exoplanet Gl 86 b and discuss the implications for
success in detecting planets around 1 − 3 Myr old T Tauri stars.
Subject headings: techniques: radial velocities -- planets and satellites: detection
-- 3 --
1.
Introduction
Since 1989, astronomers have discovered over 500 extrasolar planets, ∼92% of which
were identified using radial velocity (RV) techniques1.
improved velocity precision (∼1 m s−1 e.g., Mayor et al. 2003) have led to an abundance
of ground-based RV planet discoveries including the lightest known planet around a main
Increased time baselines and
sequence star (m sin i = 1.7 M⊕, Mayor et al. 2009) and the recent claim of a ∼3 M⊕ planet
orbiting within its star's habitable zone (Vogt et al. 2010). Ongoing exoplanet surveys
typically focus on slow-rotating FGK dwarfs. The numerous, narrow atomic lines in the
photospheres of these stars allow for precise RV determination. The SEDs of these stars also
peak in the optical region of the spectrum, facilitating their study with optical detectors, a
more mature technology than is available in other wavelength regions, such as near-infrared
(NIR).
The past few years, however, have seen a surge of interest in applying precision RV
techniques in the NIR, in part to explore M dwarfs as potential planet hosts. These late-type
stars are faint at optical wavelengths but comparatively bright at 1 − 2 µm. They are also
the most numerous stars in our Galaxy; by focusing on FGK dwarfs in most surveys to date
the majority of stars have been neglected. Also, for a given orbital period and planet mass,
the RV amplitude scales with host mass as M −2/3
∗
. Therefore, planets should be easier
to detect around lower mass M and L dwarfs. Furthermore, including the low mass stars
expands the available parameter space to include lower mass planets. The M dwarfs are
also interesting targets to the astrobiology community (Tarter et al. 2007). Because of their
lower luminosity, the habitable zones (HZ) are significantly closer to the central star than
for higher mass stars and the RV amplitudes of HZ planets are correspondingly greater,
thereby increasing the possibility of discovering a habitable Earth-mass planet (Vogt et al.
1Data from The Extrasolar Planet Encyclopedia (www.exoplanet.eu)
2010).
-- 4 --
M dwarfs may also serve as testing grounds for models of planet formation. In spite
of all the planet discoveries in the past twenty years, a unified model of planet formation
is still elusive. In the core accretion model (e.g., Pollack et al. 1996), a planetary core is
built up through the accretion of ices, dust particles, and eventually planetesimals in the
circumstellar disk. Laughlin et al. (2004) present calculations which indicate that giant
planet formation around M dwarfs under a core-accretion paradigm is inhibited as a result
of lower surface densities in the disk beyond the snowline, where Jupiter-mass worlds
are most likely to form. These planets would thus take much longer to accrete than the
typical disk lifetime; however, Kornet et al. (2006) challenge this result and argue that the
probability of planet formation actually increases with decreasing stellar mass as a result of
a more efficient particle redistribution around lower mass stars. Another avenue for giant
planet formation is the gravitational instability model (e.g., Boss 1997) wherein instabilities
in the circumstellar disk can lead to fragmentation and eventual collapse. Boss (2006)
argues that gravitational instabilities can produce Jovian worlds around low-mass stars very
rapidly. Current optical RV surveys of M dwarfs support the hypothesis that low-mass stars
are less likely to harbor planets than Sun-like stars (e.g., Endl et al. 2006; Johnson et al.
2010). Johnson et al. report that the occurrence rate of giant planets around M dwarfs
(m sin i > 0.3 MJU P , a < 2.5 AU) is 3.4+2.2
−0.9% whereas the corresponding fraction for Sun-like
stars is 7.6 ± 1.4% (Cumming et al. 2008). However, the sample size of planet-hosting M
dwarfs is small and therefore the uncertainties are large. Further exploration of the giant
planet population around M dwarfs may therefore offer some insight into which physical
processes dominate formation of these worlds.
Although the investigation of planets around M stars provides clues to their formation,
the key to learning about how planets form is to observe them in the process. Recent
-- 5 --
results on core accretion predict planet formation in millions of years (e.g., Alibert et al.
2005; Hubickyj et al. 2005; Dodson-Robinson et al. 2008). Gravitational instability models,
however, predict very fast formation times for massive planets in long period orbits: 103
(Mayer et al. 2004) to 105 (Bodenheimer 2006) years. Subsequent inward migration can
also occur on short timescales (∼105 years, Papaloizou et al. 2007) that can bring these long
period planets into the sensitivity range of RV surveys. Clearly, any data which can help to
identify this timescale will help refine, limit, and possibly eliminate some planet formation
models. Ongoing campaigns to catalog the planets around main sequence stars can not
directly provide this information. By observing late-type, pre-main sequence (PMS) T
Tauri stars, one acquires a snapshot of the early stages of planet formation around solar-like
stars.
Low-mass young stars present challenging targets for traditional RV surveys. Late
spectral types, large distances (>100 pc), and extinction from natal dust clouds make
these targets faint at optical wavelengths. They also have strong magnetic fields (e.g.,
Johns-Krull 2007) that generate large, cool star spots. These spots impact RV surveys of
young stars by introducing significant jitter which can mimic the RV modulation imposed
by a planet (Saar & Donahue 1997; Desort et al. 2007). Recent attempts at detecting
substellar companions in young stellar populations (10−100 Myr) have generally been
unsuccessful (e.g., Paulson et al. 2004; Paulson & Yelda 2006), likely because of the small
sample size and intrinsic RV variability of the targets. The youngest RV planet detected
to date is a ∼6 MJU P planet on an 850 day orbit around the 100 Myr old star HD 70573
(Setiawan et al. 2007).
Spectral line bisector analysis has historically been used to distinguish spot-induced
RV variations from companion-induced ones (e.g., Hatzes et al. 1997). Companion-induced
reflex motion does not affect the shape of an absorption line's bisector whereas a spot
-- 6 --
distorts the line symmetry. The line bisector span, the difference in bisector values at two
different heights of the line profile, is a proxy for the average slope of the line bisector.
A correlation between bisector span and RV variations suggests the RV fluctuation is
spot-induced; otherwise, it may be caused by a companion. However, there may also be
no correlation if the projected rotation velocity (v sin i) of the star is comparable to or
less than the velocity resolution of the spectrograph (Desort et al. 2007; Prato et al. 2008).
Therefore, bisector analysis is only a first step in identifying potential young planet hosts
(e.g., Hu´elamo et al. 2008; Prato et al. 2008).
A potentially more reliable method for distinguishing between spots and planets
leverages the wavelength dependence of the spot-induced RV modulation amplitude. The
reflex motion induced by a planet affects all wavelengths equally. However, the contrast
between a photosphere and a cooler star spot decreases at longer wavelengths because
of the flux-temperature scaling in the Rayleigh-Jeans limit of blackbody radiation (e.g.,
Vrba et al. 1986; Carpenter et al. 2001). As a result, the amplitude of any spot-induced RV
variability will be smaller at longer wavelengths. By observing in the visible and NIR it
may be possible to distinguish between stellar activity and the presence of a companion by
comparing the RV amplitudes at the two wavelengths. Reiners et al. (2010) point out that
the magnitude of this decrease for spotted stars is dependent on the temperature difference
between the photosphere and star spots and may not be significant if this difference is large
(i.e. ∼1000 K). Observations of T Tauri stars, however, do show reductions in RV jitter
from optical to K band wavelengths by factors of 3 − 5 (Prato et al. 2008), lending support
to the plausibility of this approach. Furthermore, the late spectral types of T Tauri stars
produce SEDs with peak emission around 1 µm and extinction of these partly embedded
sources is lessened at longer wavelengths, thus allowing for increased signal-to-noise, and
hence improved RV precision, in the NIR. Reiners et al. (2010) caution that the advantages
of NIR RV measurements may not be apparent for stars earlier than ∼M4−5. For early
-- 7 --
M stars, they assert that the decreased S/N in the optical bands is outweighed by the
number of sharp spectral features available at the shorter wavelengths. Stars later than
M5, however, do exhibit increased precision in the NIR both as a result of increased S/N in
the J and H bands as well as the appearance of FeH absorption features. Finally, T Tauri
stars allow for some relaxation in the desired RV precision. While optical surveys of main
sequence stars strive for 1 m s−1 precision or better, T Tauri stars exhibit an intrinsic NIR
RV variability of > 100 m s−1 (Prato et al. 2008; Mahmud et al. 2011). Therefore, NIR RV
precision of several tens of meters per second is more than adequate for these targets.
Thus, along with studies of planets around M and L dwarfs, RV surveys for young
planets also drive the development of improved NIR RV precision measurements. Several
groups are active in this area (e.g., Mart´ın et al. 2006; Blake et al. 2007; Hu´elamo et al.
2008; Blake et al. 2010) with some reports of NIR velocity precision comparable to that in
the optical (Bean et al. 2010; Figueira et al. 2010). While these results are encouraging,
these efforts are all focused on large (8 − 10 meter) aperture telescopes. There is therefore
a need to develop precision NIR RV techniques for smaller aperture telescopes where more
observing time is available to the community at cadences better tailored to RV surveys.
In 2004, we began the McDonald Observatory Young Planet Survey to monitor PMS
stars in the nearby (∼140 pc, Kenyon et al. 1994) Taurus-Auriga low-mass star forming
region for evidence of substellar companions. Our sample of 143 classical and weak-lined T
Tauri stars consists of V < 14 stars, most with v sin i < 20 km s−1(Herbig & Bell 1988). For
the first four years, all visible light observations were conducted using the Robert G. Tull
Coud´e Spectrometer (Tull et al. 1995) on the 2.7-meter Harlan J. Smith Telescope. In 2008,
we began to include additional observations from the Sandiford Cassegrain Spectrograph
(McCarthy et al. 1993) on the 2.1-meter Otto Struve Telescope. Observations at the Kitt
Peak National Observatory (KPNO) 4-meter Mayall Telescope using the cassegrain echelle
-- 8 --
spectrograph were also initiated in late 2008. Promising targets from these observations,
based on measured RV variations and the results of bisector analysis, are selected for follow
up K band observations at the 3-meter NASA Infrared Telescope Facility (IRTF) using the
high-resolution Cassegrain-mounted echelle spectrograph, CSHELL (Tokunaga et al. 1990;
Greene et al. 1993). Given the 1−3 Myr ages of our targets, a positive detection of a young
exoplanet will provide a unique datapoint for giant planet formation theory.
In this paper, we discuss our methodology for measuring precise NIR RVs with
CSHELL. In §2 we discuss our observing strategy and data reduction algorithms. In §3
we present our methodology for using telluric absorption features to measure the RVs of
our targets. We present results from observations of an RV standard star and a known
exoplanet in §4, and we discuss the limitations and possibilities of our technique in §5.
2. Observation strategy and data reduction
Observations were taken at the 3 meter NASA Infrared Telescope Facility (IRTF)
using CSHELL (Tokunaga et al. 1990; Greene et al. 1993). CSHELL is a long-slit echelle
spectrograph (1.08 µm − 5.5 µm) that uses a Circular Variable Filter (CVF) to isolate a
single order onto a 256×256 InSb detector array. We used the CVF to isolate a 50 A segment
of spectrum centered at 2.298 µm. This region contains numerous deep photospheric
absorption lines from the CO ν = 2−0 bandhead as well a rich set of predominately CH4
telluric absorption features which we use as a wavelength reference. The 0′′.5 slit yielded a
typical FWHM of 2.6 pixels (∼0.5 A, measured from arc lamp spectra) corresponding to a
spectral resolving power of R ∼ 46,000.
We observed RV standards, a known exoplanet candidate, and several T Tauri planet
host candidates in February 2008 (8 nights), November 2008 (6 nights), February 2009 (2
-- 9 --
nights), November 2009 (5 nights), and February 2010 (8 nights). At the beginning of
each night, we imaged twenty flat fields, each with a 20 second integration time, using a
continuum lamp to illuminate the entire slit. We also imaged the same number of 20-second
dark frames. Additionally, we imaged six Ar-Kr-Xe emission lines, changing the CVF while
maintaining the grating position, to determine the wavelength reference. All of our target
data were obtained using 10′′ nodded pairs to enable subtraction of sky emission, dark
current, and detector bias. Integration times for each nod were typically 600 seconds; for
fainter targets we took multiple contiguous nod pairs. The signal-to-noise ratio (S/N) for
all targets varied significantly depending on cloud cover, seeing, guiding errors, etc., with
typical values for all targets of ∼50/pixel per nod position.
Our data reduction strategy closely follows that described in Johns-Krull et al. (1999)
and was implemented entirely in IDL. We first produced a nightly master dark frame by
averaging the twenty individual dark exposures and found a median dark current of 0.24
e−/s. We produced a nightly normalized flat field by averaging the flat field exposures,
subtracting the master dark frame, and then dividing by the mean of the dark-subtracted
master flat. For our target data, we first subtracted the nod pairs to create a difference
image. The difference image was subsequently divided by the normalized flat field. We
estimated the read noise from the standard deviation of a Gaussian fit to a histogram of
the pixel values in the difference image (∼30 e−). We then identified the location of the
curved spectral traces in the difference image by fitting a second-order polynomial to the
location of maximum and minimum flux along the dispersion direction.
To optimally extract the spectrum, we divided each nod pair into four equally spaced
bins of 64 columns along the dispersion direction. Within each of these 64 column bins
we constructed a 10× oversampled "slit function" (i.e., the distribution of flux in the
cross-dispersion direction). First, a rough estimate of the spectrum was created by summing
-- 10 --
the pixels in each column of the difference image for each nod position. The limits included
in this sum are from the midpoint between the two nod positions on the detector to the
edge of the area that is well illuminated by the flat lamp. This is typically 60-70 pixels in
each column for each nod position. Next, each pixel in the bin was sorted by its distance
from the order center for the column the given pixel falls in. The flux in each pixel was
divided by the rough estimate of the spectrum for its appropriate column to normalize all
the pixels going into the slit function estimate. The flux in these offset ordered pixels was
then median filtered with a 7 point moving box. A flux estimate for each oversampled pixel
was then determined by taking the median of all the pixels that fell in a given subpixel.
This then formed our oversampled master slit function. The multiple median filters
generally remove the effects of cosmic rays and uncorrected bad pixels on the determination
of the slit function. We then fit this master slit function to a three Gaussian model: a
central Gaussian flanked by two satellite Gaussians. The amplitude, center, and width of
each Gaussian were fit as free parameters using the IDL implementation of the AMOEBA
non-linear least-squares (NLLS) fitting algorithm (Nelder & Mead 1965). The resulting
model was then normalized to unit area. This algorithm produces four model slit functions,
one for each bin. However, the actual slit function is a smoothly varying function of column
number. Therefore, to smooth out the transitions from bin to bin, we create 256 column
slit functions by linearly interpolating between the four bin slit functions.
To determine the total flux in each column of the spectrum, we calculated the scale
factor that best matches our model slit function to the column data, per the recipe
described in Horne (1986). In order to mask out spurious flux levels from cosmic rays, we
implemented an iterative sigma-clipping algorithm. This algorithm starts with an estimate
of the total noise from the measured read noise in the differenced image plus the Poisson
noise from the target. We then subtracted our model fit from the data in each detector
column and masked those pixels whose residual was 3-σ greater than the estimated noise.
-- 11 --
We iterated through this process 1−2 times until the scale factor converged, thus providing
an optimal value of the spectrum in that column that is largely immune to hot pixels,
cosmic rays, etc. This algorithm also provides an estimate of the flux uncertainty at each
location along the spectrum.
3. Radial velocity determination
The telluric absorption features in the K-band provide an absolute wavelength and
instrumental profile reference, similar in concept to the iodine gas cell technique used in
high-precision optical RV exoplanet surveys (Butler et al. 1996). Using the atmosphere as
a "gas cell" lets us superimpose onto our spectra a relatively stable wavelength reference
which follows the same optical path as the light from the science target. This helps alleviate
uncertainties introduced by variable slit illumination, changing optical path lengths, etc.
We determined the radial velocities of our targets using a spectral modeling technique
similar to the one presented in Blake et al. (2007), Prato et al. (2008), and Figueira et al.
(2010). We modeled the stellar spectrum and the telluric features using two high-resolution
template spectra. For the stellar spectrum, we employed NextGen stellar atmosphere
models (Hauschildt et al. 1999) tailored to the Tef f , log g, and metallicity of our targets.
We used SYNTHMAG (Piskunov 1999) to generate spectra from the NextGen models along
with atomic (Kupka et al. 2000) and CO (Goorvitch 1994) line lists. These spectra are
sampled at a resolution ∼14 times greater than our observations. We modeled the telluric
features using the NOAO telluric absorption spectrum (Livingston & Wallace 1991) which
has a resolution ∼4 times higher than that of our observations.
In order to match the model to our observations, we fit a velocity shift and a power law
scaling factor for each template separately. The power law scaling accounts for differences
in line strength between the observed and template spectra (e.g., McCullough et al. 2006).
-- 12 --
The stellar spectrum was then interpolated onto the wavelength scale for the telluric
spectrum. To match stars with measurable rotation, we fit for the value of v sin i that
optimally broadens the stellar template spectrum using a disk-integration routine with
a limb-darkening coefficient of 0.2 (based on fitting the model photosphere to a linear
limb-darkening law). For the main sequence stars observed for this paper, however,
we fixed v sin i to zero since it is not meaningfully detected above the instrumental
broadening. Because the normalized telluric spectrum models the relative transmission of
the atmosphere, we multiplied the telluric spectrum and broadened stellar spectrum together
to create a composite spectrum and convolved the result with a Gaussian instrumental
profile (IP) to model the instrument response. We experimented with multi-Gaussian
models of the IP (e.g., Valenti et al. 1995; Bean et al. 2007) but found no improvement in
the RV precision. We then fit the continuum level with a second-order polynomial. Finally,
we binned the 4× oversampled composite model down to the resolution of our CSHELL
data. The resultant model has nine free parameters (two velocity shifts, two scaling factors,
v sin i, IP FWHM, and three continuum coefficients). To simultaneously determine the best
fit to these parameters, we again used the IDL implementation of the AMOEBA NLLS
algorithm.
Because CSHELL is cassegrain-mounted, instrument flexure results in a wavelength
dispersion on the detector that changes with each observation. We determined an initial
dispersion solution using the Ar-Kr-Xe lamp lines imaged at the start of each night (§2).
The six individual arc lamp spectral images were stacked to create a composite spectrum.
The combined arc spectrum was then extracted using the order tracing from that night's
highest S/N stellar spectrum. We found an initial dispersion solution by fitting a third
order polynomial to the locations of these emission lines. This initial polynomial serves as
a starting point to refine the dispersion solution. For each target observation, we take the
observed spectrum and the best-fit model and divide both into eight, 6.25 A wide, bins.
-- 13 --
This breaks up the observed and model spectra in to eight pieces. We then take each piece
of observed spectrum and cross-correlate it with the corresponding piece of model spectrum
to measure the pixel shift between the two. The pixel shift as a function of spectrum bin
provides a crude estimate of how much the observed dispersion solution differs from the
initial solution across the detector. The eight pixel shift values are fit to a second-order
polynomial. By interpolating this polynomial across all detector columns, we estimate by
how much the wavelength has shifted per pixel relative to the initial dispersion. A new
wavelength solution is determined by adding this shift to the initial solution. We then
produce a new model fit using the refined wavelength solution. We iterate this algorithm
∼5 times until the wavelength solution converges.
Figure 1 illustrates the outcome of this modeling for one observation of our RV
standard, GJ 281. Radial velocities for each observation were calculated by taking the
difference in the velocity shifts between the best-fit stellar and telluric spectra. These
velocities were in turn corrected for barycentric motion determined for the mid-time of each
exposure 2.
For each observation, we estimated uncertainties introduced by the photon-limited
errors of our observations. These uncertainties in turn have two main components: the
first is from the information content in the stellar spectrum itself and the second is in the
information content in the telluric spectrum which is used as a wavelength calibrator. As
described in Butler et al. (1996), the intrinsic RV error of a spectrum is the quadrature sum
of these two error contributions. This in turn is a function of the slope of the intensity with
respect to wavelength at each pixel and the S/N. For a given S/N, a spectrum with more
numerous and sharper absorption features will have more information content than one with
2Barycentric corrections were calculated using helcorr.pro, an IDL routine based on
the IRAF task noao.astutils.rvcorrect.
-- 14 --
fewer, broader features. We used the best-fit model photosphere to calculate the intensity
derivative and the measured flux uncertainty (estimated from the measured read noise plus
the target Poisson noise, §2) to determine the S/N for each pixel. The combined errors from
each pixel determine the photon-limited uncertainty on the final RV. We derived the errors
introduced by the telluric spectrum in exactly the same manner, replacing the photosphere
model with the telluric one. The final uncertainty on the velocity was calculated by adding
the stellar and telluric errors in quadrature.
A final, nightly RV was determined by calculating the average of the individual nod
RVs, weighted by the median S/N across each spectrum. The uncertainties were computed
by taking the weighted standard deviation of the nod RVs and dividing by the square root
of the number of nods.
4. Results
4.1. GJ 281
To determine the precision of our approach, we obtained 28 observations of GJ
281 over the past two years. This is a late-type star (M0, K mag = 5.9, Tef f = 3776
K, Casagrande et al. 2008) known to have a stable RV (r.m.s ∼ 12 m s−1, Endl et al.
2003). We determined the nightly RVs and uncertainties using the approach described in
Section 3 (see Figure 2 and Table 1). The median uncertainty from photon statistics is
40 m s−1 per nod with the telluric uncertainty ∼50% greater than the stellar spectrum
uncertainty. The standard deviation of the mean within a sequence of contiguous nod pairs
(σRV /√N) is typically 37 m s−1. After averaging, the standard deviation of the nightly
RVs over the entire 24 month observing window is ∼58 m s−1. Since this is greater than
the variability seen in a given night, we need to estimate what additional systematic error
-- 15 --
we are introducing between nights. Assuming that the night-to-night systematics add in
quadrature with the variability over multiple contiguous nods to give us our overall 58
m s−1 standard deviation, (σ2
nightly), we estimate √582 − 372 = 45 m s−1 as
T OT = σ2
nods + σ2
our night-to-night systematic limit.
To better understand the systematic errors, we looked for differences in the measured
velocities as a function of nod position. Subtracting the A beam velocities from the B
beam velocities, we found a mean difference between the beams of 66 m s−1. The standard
deviation of these differences was 131 m s−1, while the standard deviation of the mean was
22 m s−1. The standard deviation of the mean of just the A beam velocities was 15 m s−1;
for the B beam velocities, this was 17 m s−1. This suggests that there is a statistically
significant difference between the velocities at the two nod positions which gets averaged
out at some level when combining all observations in a single night. A simple subtraction
from the A velocities to remedy this offset is not sufficient, however. In a few nod pairs,
the B velocities are higher than the A velocities and differencing makes this offset worse.
Because of this, the variability seen in the binned velocities is nearly identical whether or
not the mean offset is removed.
It is not clear why there is a velocity offset between the beams. While we expect to see
different distortions in the two nod positions, the use of telluric features as a common-path
calibrator should account for this. One concern with our observing setup is the asymmetric
distribution of telluric lines across the spectrum. The long wavelength half of our spectra is
much richer in telluric features than the short-wavelength half. It is therefore possible that
we are creating false higher order terms as we track the change in the dispersion across the
slit. We tested this by recalculating our RV determination for all the GJ 281 observations
using only the left half of the spectrum and then using only the right half of the spectrum.
Fitting only the left half results in a mean difference between the A and B beams of 33
-- 16 --
m s−1 whereas restricting ourselves to the telluric-rich right half results in a mean difference
between the beams of 10−6 m s−1. While this would seem to support the hypothesis that
an asymmetric telluric line distribution may be introducing systematic errors, the standard
deviation of the mean in each beam for both cases is large enough to make the statistical
significance of these results unimportant. Interpretation of the much lower RV scatter
in the right half of the spectrum is therefore unclear. It is likely that we simply do not
have enough signal to noise to effectively separate out this error from our other sources of
uncertainty. Future observations of very bright targets may help identify the sources of our
systematic error. We note that the existence of a systematic shift between the two beams
combined with our use of the standard deviation of the measured RVs in both nod positions
will most likely over estimate the uncertainty in our velocity measurements.
4.2. Gl 86
In addition to RV standards and T Tauri planet host candidates, we also obtained six
observations of the known exoplanet host Gl 86 (K mag = 4.1, Tef f = 5350 K, log g = 4.6,
Flynn & Morell 1997) in November 2008. Queloz et al. (2000) presented evidence for a hot
Jupiter around Gl 86 with m sin i = 4.0 MJU P and P = 15.78 days. Figueira et al. (2010)
used Gl 86 as a test case for their precision NIR RV work with the CRIRES spectrograph
on the VLT and were able to show that their observations were consistent with the
Queloz et al. orbital solution, allowing for errors in time of periastron (T0) determination
and drift in the center-of-mass velocity (Vr) as a result of the presence of a long period
companion (Eggenberger et al. 2003). We present our nightly RVs for Gl 86 in Table 2
and Figure 3. The median uncertainty from photon statistics is 41 m s−1 with the telluric
uncertainty ∼43% greater than the stellar spectrum uncertainty. The standard deviation of
the mean within a sequence of contiguous nod pairs is typically 58 m s−1. Figure 3 plots our
-- 17 --
binned RVs with those from Figueira et al. and an orbital fit to both datasets. Our error
bars are determined by adding the results from our binning algorithm (§3) in quadrature
with the 45 m s−1 systematic uncertainty determined from our RV standard (§4.1). When
fitting an orbit model to the data, we allowed T0 and Vr to be fit as free parameters, while
fixing all other values to those from Queloz et al. (2000). We show good agreement with
previously published values exhibiting a ∼46 m s−1 standard deviation in the residuals.
All of our RVs differ from the orbital solution by no more than 0.5 sigma; the
mean difference is ∼0.24 sigma. Assuming normally distributed errors, we would expect
1 − 2 of these residuals to be greater than one sigma. To explore this further, we
assumed that the underlying uncertainties are gaussian and drawn from a distribution
with σ = √582 + 452 = 73 m s−1. The variance in the residuals is dominated by the first
RV measurement; excluding that point we find a standard deviation in the residuals of
16 m s−1. We used a Monte Carlo analysis to estimate that the probability of having at
least five out of six points within 16 m s−1 of the mean, given σ = 73 m s−1, is ∼0.08%.
We further explored the false alarm probability (FAP) that the phase coherence from our
observations is the result of chance. We performed a Monte Carlo analysis by holding the
times fixed and sampling the CSHELL velocities (and corresponding uncertainties) with
replacement over 104 iterations. For each iteration, we fit an orbit model as described above
and recorded how often the reduced χ2 of the fit was less than that of the fit to the observed
velocities. We find that it is highly unlikely that the phase coherence is a result of chance,
with a FAP of 0.0003. We are therefore very confident that our RVs are consistent with the
known planetary-mass object orbiting Gl 86. The inconsistency between our estimated error
bars and the residual scatter is suspicious, however. The scatter in the residuals is more
comparable to the estimated photon uncertainty. Our 45 m s−1 systematic uncertainty is
measured over a period of two years whereas the RVs for Gl 86 were measured over a single
run. If the systematics were "quiet" on this run, then 45 m s−1 may not be the appropriate
-- 18 --
value to use. We see some evidence for the appearance of "quiet" times in our data on GJ
281. During the November 2008 run the standard deviation of our nightly (6 nights) GJ
281 data is 39 m s−1, well below the 58 m s−1 of the entire data set. This is the same run on
which the Gl 86 data were obtained. We therefore believe that our uncertainties are most
likely overestimates.
5. Discussion
NIR spectroscopy has only recently begun to play a role in the search for substellar
companions. Mart´ın et al. (2006) combined optical and H band observations to look for
giant planets around the young brown dwarf LP 944-20; they concluded that the observed
optical RV modulations were driven by inhomogeneous surface features (i.e. clouds).
Blake et al. (2007) used high resolution K-band spectroscopy to investigate the presence of
giant planets around a population of L dwarfs, achieving a precision of 300 m s−1. They
find no evidence for companions with M sin i > 2MJ and P < 3 days in their sample of
nine targets. Setiawan et al. (2008) reported the detection of a ∼10 MJ planet on a 3.5
day orbit around the 10 Myr old star TW Hydra. However, H band observations presented
by Hu´elamo et al. (2008) reveal a strong wavelength dependence in the velocity amplitude,
thus casting doubt on the presence of a companion and suggesting that spots are the
cause of TW Hydra's RV variations. Prato et al. (2008) observed two potential planet-host
candidates in the NIR, selected on the basis of their optical variability. To within their
measurement precision, they detect no K band RV modulation, implying that spots are
responsible for the apparent optical RV variability.
The aforementioned interest in targeting M dwarfs as hosts of habitable planets (§1)
has spurred vast improvements in NIR precision, bringing it closer to that of optical surveys.
Bean et al. (2010) report achieving a long term precision in the H band of ∼5 m s−1 on late
-- 19 --
M dwarfs using CRIRES at the VLT with the aid of an ammonia gas cell. Figueira et al.
(2010) report a comparable precision also on CRIRES but using telluric absorption features
as a wavelength reference. Based on six years of NIRSPEC data, Blake et al. (2010) report
a precision of 50 m s−1 for their sample of K dwarfs and 200 m s−1 for L dwarfs also through
the use of telluric lines.
In an earlier paper from our survey of young, low-mass stars, Prato et al. (2008)
presented evidence that the RV variability of two T Tauri stars, DN Tau and V836 Tau, is
the result of star spots. These two targets exhibited an optical RV standard deviation of
438 m s−1 and 742 m s−1, respectively. K band RVs, however, showed a standard deviation
of 144 m s−1 for DN Tau and 149 m s−1 for V836 Tau. These values are consistent with
the measurement uncertainties of ∼165 m s−1 and ∼290 m s−1, respectively. These stars
therefore exhibit no evidence of RV jitter in the K band. The undetected NIR RV variability
demonstrates two important points. First, optical RV surveys can not rely on bisector
analysis alone to ascertain the mechanism for RV variability. Both DN Tau and V836 Tau
showed significant periodic optical variability and neither exhibited a correlation between
RV and bisector span. Multiwavelength observations are therefore an essential element for
testing the planet hypothesis, especially in young stars. Second, they support the hypothesis
that RV jitter from stellar activity should be significantly lower at longer wavelengths. We
have demonstrated ∼58 m s−1 precision with CSHELL over two years and the ability to
detect a known 4 MJU P planet. If the ∼150 m s−1 variability observed in DN Tau and
V836 Tau is typical of T Tauris, our 58 m s−1 precision is well below the intrinsic noise floor
of our targets suggesting that our methodology is well-suited to detecting "hot Jupiters"
around young stars. However, we caution against extrapolating from two young stars to
the NIR behavior of other T Tauri stars. Mahmud et al. (2011) present another target
from our Taurus survey that exhibits a ∼430 m s−1 modulation in the K band, which we
believe to be caused by spots, suggesting that the NIR RV behavior of T Tauris may vary
-- 20 --
significantly from star to star.
The authors thank R. White and J. Bailey for productive discussions on data reduction
and our anonymous referee for offering many useful suggestions which improved the
manuscript. We acknowledge the SIM Young Planets Key Project for research support;
funding was also provided by NASA Origins Grants 05-SSO05-86 and 07-SSO07-86. This
work made use of the SIMBAD database, the NASA Astrophysics Data System, and the
Two Micron All Sky Survey (2MASS), a joint project of the University of Massachusetts
and IPAC/Caltech, funded by NASA and the NSF. We recognize the significant cultural
role that Mauna Kea plays in the indigenous Hawaiian community and are grateful for the
opportunity to observe there.
Facilities: IRTF (CSHELL).
-- 21 --
REFERENCES
Alibert, Y., Mordasini, C., Benz, W., & Winisdoerffer, C. 2005, A&A, 434, 343
Bean, J. L., McArthur, B. E., Benedict, G. F., Harrison, T. E., Bizyaev, D., Nelan, E., &
Smith, V. V. 2007, AJ, 134, 749
Bean, J. L., Seifahrt, A., Hartman, H., Nilsson, H., Wiedemann, G., Reiners, A., Dreizler,
S., & Henry, T. J. 2010, ApJ, 713, 410
Blake, C. H., Charbonneau, D., & White, R. J. 2010, ApJ, 723, 684
Blake, C. H., Charbonneau, D., White, R. J., Marley, M. S., & Saumon, D. 2007, ApJ, 666,
1198
Bodenheimer, P. 2006, Historical notes on planet formation (Cambridge University Press),
1
Boss, A. P. 1997, Science, 276, 1836
-- . 2006, ApJ, 643, 501
Butler, R. P., Marcy, G. W., Williams, E., McCarthy, C., Dosanjh, P., & Vogt, S. S. 1996,
PASP, 108, 500
Carpenter, J. M., Hillenbrand, L. A., & Skrutskie, M. F. 2001, AJ, 121, 3160
Casagrande, L., Flynn, C., & Bessell, M. 2008, MNRAS, 389, 585
Cumming, A., Butler, R. P., Marcy, G. W., Vogt, S. S., Wright, J. T., & Fischer, D. A.
2008, The Publications of the Astronomical Society of the Pacific, 120, 531
Desort, M., Lagrange, A.-M., Galland, F., Udry, S., & Mayor, M. 2007, A&A, 473, 983
-- 22 --
Dodson-Robinson, S. E., Bodenheimer, P., Laughlin, G., Willacy, K., Turner, N. J., &
Beichman, C. A. 2008, ApJ, 688, L99
Eggenberger, A., Udry, S., & Mayor, M. 2003, in Astronomical Society of the Pacific
Conference Series, Vol. 294, Scientific Frontiers in Research on Extrasolar Planets,
ed. D. Deming & S. Seager, 43 -- 46
Endl, M., Cochran, W. D., Kurster, M., Paulson, D. B., Wittenmyer, R. A., MacQueen,
P. J., & Tull, R. G. 2006, ApJ, 649, 436
Endl, M., Cochran, W. D., Tull, R. G., & MacQueen, P. J. 2003, AJ, 126, 3099
Figueira, P., et al. 2010, A&A, 511, 55
Flynn, C., & Morell, O. 1997, MNRAS, 286, 617
Goorvitch, D. 1994, ApJSS, 95, 535
Greene, T. P., Tokunaga, A. T., Toomey, D. W., & Carr, J. B. 1993, Proc. SPIE, 1946, 313
Hatzes, A. P., Cochran, W. D., & Johns-Krull, C. M. 1997, ApJ, 478, 374
Hauschildt, P. H., Allard, F., & Baron, E. 1999, ApJ, 512, 377
Herbig, G. H., & Bell, K. R. 1988, Third Catalog of Emission-Line Stars of the Orion
Population (Lick Observatory Bulletin)
Horne, K. 1986, PASP, 98, 609
Hubickyj, O., Bodenheimer, P., & Lissauer, J. J. 2005, Icarus, 179, 415
Hu´elamo, N., et al. 2008, A&A, 489, L9
Johns-Krull, C. M. 2007, ApJ, 664, 975
-- 23 --
Johns-Krull, C. M., Valenti, J. A., & Koresko, C. 1999, ApJ, 516, 900
Johnson, J. A., et al. 2010, The Publications of the Astronomical Society of the Pacific,
122, 149
Kenyon, S. J., Dobrzycka, D., & Hartmann, L. 1994, AJ, 108, 1872
Kornet, K., Wolf, S., & R´ozyczka, M. 2006, A&A, 458, 661
Kupka, F. G., Ryabchikova, T. A., Piskunov, N. E., Stempels, H. C., & Weiss, W. W. 2000,
Baltic Astronomy, 9, 590
Laughlin, G., Bodenheimer, P., & Adams, F. C. 2004, ApJ, 612, L73
Livingston, W., & Wallace, L. 1991, An atlas of the solar spectrum in the infrared from
1850 to 9000 cm-1 (1.1 to 5.4 micrometer) (National Solar Observatory)
Mahmud, N., Crockett, C., Prato, L., Johns-Krull, C. M., Huerta, M., Hartigan, P., Jaffe,
D. T., & Beichman, C. A. 2011, ApJ, submitted
Mart´ın, E. L., Guenther, E., Osorio, M. R. Z., Bouy, H., & Wainscoat, R. 2006, ApJ, 644,
L75
Mayer, L., Quinn, T., Wadsley, J., & Stadel, J. 2004, ApJ, 609, 1045
Mayor, M., et al. 2003, The Messenger (ISSN0722-6691), 114, 20
-- . 2009, A&A, 507, 487
McCarthy, J. K., Sandiford, B. A., Boyd, D., & Booth, J. 1993, Astronomical Society of the
Pacific, 105, 881
McCullough, P. R., et al. 2006, ApJ, 648, 1228
Nelder, J., & Mead, R. 1965, Computer Journal, 7, 308
-- 24 --
Papaloizou, J. C. B., Nelson, R. P., Kley, W., Masset, F. S., & Artymowicz, P. 2007,
Protostars and Planets V, 655
Paulson, D. B., Cochran, W. D., & Hatzes, A. P. 2004, AJ, 127, 3579
Paulson, D. B., & Yelda, S. 2006, PASP, 118, 706
Piskunov, N. 1999, in Astrophysics and Space Science Library, Vol. 243, Polarization, ed.
K. N. Nagendra & J. O. Stenflo, 515
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., & Greenzweig,
Y. 1996, Icarus, 124, 62
Prato, L., Huerta, M., Johns-Krull, C. M., Mahmud, N., Jaffe, D. T., & Hartigan, P. 2008,
ApJ, 687, L103
Queloz, D., et al. 2000, A&A, 354, 99
Reiners, A., Bean, J. L., Huber, K. F., Dreizler, S., Seifahrt, A., & Czesla, S. 2010, ApJ,
710, 432
Saar, S. H., & Donahue, R. A. 1997, ApJ, 485, 319
Setiawan, J., Henning, T., Launhardt, R., Muller, A., Weise, P., & Kurster, M. 2008,
Nature, 451, 38
Setiawan, J., Weise, P., Henning, T., Launhardt, R., Muller, A., & Rodmann, J. 2007, ApJ,
660, L145
Tarter, J. C., et al. 2007, Astrobiology, 7, 30
Tokunaga, A. T., Toomey, D. W., Carr, J., Hall, D. N. B., & Epps, H. W. 1990, Proc.
SPIE, 1235, 131
-- 25 --
Tull, R. G., MacQueen, P. J., Sneden, C., & Lambert, D. L. 1995, PASP, 107, 251
Valenti, J. A., Butler, R. P., & Marcy, G. W. 1995, PASP, 107, 966
Vogt, S. S., Butler, R. P., Rivera, E. J., Haghighipour, N., Henry, G. W., & Williamson,
M. H. 2010, arXiv, astro-ph.EP
Vrba, F. J., Rydgren, A. E., Chugainov, P. F., Shakovskaia, N. I., & Zak, D. S. 1986, ApJ,
306, 199
This manuscript was prepared with the AAS LATEX macros v5.2.
-- 26 --
t
n
a
t
s
n
o
c
+
x
u
l
f
d
e
z
i
l
a
m
r
o
N
1.5
1.0
0.5
0.0
2.295
2.296
2.297
2.298
Wavelength (microns)
(a)
(b)
(c)
(d)
2.299
Fig. 1. -- Illustration of NIR spectrum modeling for our highest S/N observation of GJ
281 (JD 2455252.906): (a) NextGen photosphere model, (b) telluric template, (c) combined
model (solid line) with CSHELL observation (dots), (d) residuals. All plots are normalized
to the final model and have constants added for visual clarity.
-- 27 --
Feb 2008
Nov 2008
Feb 2009
Nov 2009
Feb 2010
)
1
−
s
m
k
(
V
R
c
i
r
t
n
e
c
o
i
l
e
H
20.2
20.1
20.0
19.9
19.8
1
3
5
7
271
273
275
277
375
376
Days since Feb 13, 2008
641
643
645
647
649
651
726
728
730
732
734
Fig. 2. -- Binned K band radial velocities of GJ 281 from Feb 2008 − Feb 2010. The error
bars are the standard deviation of the mean of all nods obtained on that night (σnods/√N );
they exhibit a median value of ∼37 m s−1. The standard deviation of the binned RVs over
the entire observing period is 58 m s−1.
-- 28 --
)
1
−
s
m
k
(
V
R
e
v
i
t
l
a
e
R
0.4
0.2
0.0
−0.2
−0.4
0.0
0.2
0.4
0.6
Relative Phase
0.8
1.0
Fig. 3. -- Known exoplanet Gl 86 b. Open blue squares are NIR RVs measured using CRIRES
(Figueira et al. 2010). Filled red circles are our binned K band RVs from November 2008.
Black line is orbit model fit to the combined datasets. Weighted standard deviation of our
residuals is 46 m s−1. The bar in the lower left indicates the median uncertainty from photon
statistics.
-- 29 --
Table 1. GJ 281 Radial Velocities
JD - 2450000 Heliocentric RV (km s−1)
σnods/√N (m s−1)
4510.914
4511.910
4512.914
4513.895
4514.910
4515.902
4516.902
4781.137
4783.125
4784.109
4785.082
4786.109
4787.133
4884.906
4885.867
5151.137
5153.160
5156.148
5158.152
5160.145
5235.934
20.096
19.985
19.981
20.089
20.092
20.054
20.129
20.108
20.134
20.161
20.114
20.048
20.138
20.065
20.126
20.185
20.137
20.094
20.054
20.161
19.988
70.8
14.2
67.7
37.4
60.6
5.5
51.6
12.3
30.7
26.0
38.4
71.0
59.0
28.0
29.3
41.8
50.9
6.4
30.7
19.0
42.7
-- 30 --
Table 1 -- Continued
JD - 2450000 Heliocentric RV (km s−1)
σnods/√N (m s−1)
5236.922
5237.922
5238.926
5239.918
5240.926
5241.906
5242.902
19.996
20.171
20.135
20.053
20.150
20.107
20.066
32.8
26.3
10.8
61.0
63.8
12.8
86.6
-- 31 --
Table 2. Gl 86 Radial Velocities
JD - 2450000 Heliocentric RV (km s−1)
σnods/√N (m s−1)
4781.902
4782.891
4783.879
4784.883
4785.891
4786.891
56.303
56.087
56.008
55.974
56.054
56.118
199.5
73.0
52.8
52.3
65.4
81.2
|
1504.01543 | 1 | 1504 | 2015-04-07T10:37:58 | A Search for Subkilometer-sized Ordinary Chondrite Like Asteroids in the Main-Belt | [
"astro-ph.EP"
] | The size-dependent effects of asteroids on surface regolith and collisional lifetimes suggest that small asteroids are younger than large asteroids. In this study, we performed multicolor main-belt asteroid (MBA) survey by Subaru telescope/Suprime-Cam to search for subkilometer-sized ordinary chondrite (Q-type) like MBAs. The total survey area was 1.5 deg^2 near ecliptic plane and close to the opposition. We detected 150 MBAs with 4 bands (B, V , R, I) in this survey. The range of absolute magnitude of detected asteroids was between 13 and 22 magnitude, which is equivalent to the size range of kilometer to sub-kilometer diameter in MBAs. From this observation, 75 of 150 MBAs with color uncertainty less than 0.1 were used in the spectral type analysis, and two possible Q-type aster- oids were detected. This mean that the Q-type to S-type ratio in MBAs is < 0.05. Meanwhile, the Q/S ratio in near Earth asteroids (NEAs) has been estimated to be 0.5 to 2 (Binzel et al., 2004; Dandy et al., 2003). Therefore, Q-type NEAs might be delivered from the main belt region with weathered, S-type surface into near Earth region and then obtain their Q-type, non- weathered surface after undergoing re-surfacing process there. The resur- facing mechanisms could be: 1. dispersal of surface material by tidal effect during planetary encounters (Binzel et al., 2010; Nesvorny et al., 2010), 2. the YORP spin-up induced rotational-fission (Polishook et al., 2014) or surface re-arrangement, or 3. thermal degradation (Delbo et al., 2014). | astro-ph.EP | astro-ph |
A Search for Subkilometer-sized Ordinary Chondrite
Like Asteroids in the Main-Belt
H. W. Lina, Fumi Yoshidab, Y. T. Chenc, W. H. Ipa,d, C. K. Changa
aInstitute of Astronomy, National Central University, Taoyuan 32001, Taiwan
bNational Astronomical Observatory of Japan, 2-21-1 Osawa, Mitaka, Tokyo 181-8588,
cInstitute of Astronomy and Astrophysics, Academia Sinica, P. O. Box 23-141, Taipei
dSpace Science Institute, Macau University of Science and Technology, Taipa, Macau
106, Taiwan
JAPAN
Abstract
The size-dependent effects of asteroids on surface regolith and collisional
lifetimes suggest that small asteroids are younger than large asteroids. In this
study, we performed multicolor main-belt asteroid (MBA) survey by Subaru
telescope/Suprime-Cam to search for subkilometer-sized ordinary chondrite
(Q-type) like MBAs. The total survey area was 1.5 deg2 near ecliptic plane
and close to the opposition. We detected 150 MBAs with 4 bands (B, V ,
R, I) in this survey. The range of absolute magnitude of detected asteroids
was between 13 and 22 magnitude, which is equivalent to the size range of
kilometer to sub-kilometer diameter in MBAs.
From this observation, 75 of 150 MBAs with color uncertainty less than
0.1 were used in the spectral type analysis, and two possible Q-type aster-
oids were detected. This mean that the Q-type to S-type ratio in MBAs is
< 0.05. Meanwhile, the Q/S ratio in near Earth asteroids (NEAs) has been
estimated to be 0.5 to 2 (Binzel et al., 2004; Dandy et al., 2003). Therefore,
Q-type NEAs might be delivered from the main belt region with weathered,
S-type surface into near Earth region and then obtain their Q-type, non-
weathered surface after undergoing re-surfacing process there. The resur-
facing mechanisms could be: 1. dispersal of surface material by tidal effect
during planetary encounters (Binzel et al., 2010; Nesvorn´y et al., 2010), 2.
the YORP spin-up induced rotational-fission (Polishook et al., 2014) or sur-
face re-arrangement, or 3. thermal degradation (Delbo et al., 2014).
Keywords:
Preprint submitted to Icarus
September 24, 2018
Asteroids, surfaces, Asteroids, composition, Asteroids, Regoliths
1. Introduction
The taxonomic type of asteroid has been studied extensively to under-
stand the mineral composition of asteroids.
It is mostly based on the as-
teroid's colors and spectra in optical wavelength. Numerous types (such as
S, C, D, B and V) have been identified (Bowell et al., 1978; Tholen, 1984;
Zellner et al., 1985; Bus and Binzel, 2002a,b; DeMeo et al., 2009; DeMeo and Carry,
2013, 2014). Space weathering effects and related color-spectrum correla-
tions for the main-belt asteroids (MBAs) and near-Earth asteroids (NEAs)
have also been studied (Chapman, 1996; Binzel et al., 2001; Chapman, 2004;
Clark et al., 2001; Jedicke et al., 2004; Nesvorn´y et al., 2005; Willman et al.,
2008, 2010; Willman and Jedicke, 2011; Thomas et al., 2012). These studies
have been primarily based on relatively larger asteroids; only few studies
done for kilometer to sub-kilometer asteroids because of the requirement of
large telescopes to determine their colors or spectra, have been conducted.
Asteroids with sizes below the kilometer range are most likely collisional
fragments of large asteroids (Davis et al., 2002; Morbidelli et al., 2009). There-
fore, their surfaces should have lower degree of space weathering compared
with larger asteroids, which have survived throughout the history of the solar
system (Binzel et al., 2001, 2004; Bus and Binzel, 2002b). Some small, sev-
eral hundred meter sized NEAs observed in detail while during close approach
to the Earth, showed Q-type spectra which are similar to those of the ordi-
nary chondrite (OC) with low degree of space weathering (Tholen, 1984). Re-
searchers also reported that a spectral transition could occur between S-type
and Q-type asteroids (Binzel et al., 1996, 2004; Dandy et al., 2003). These
results indicated that S-type asteroids are likely Q-type asteroids, with their
surface materials originally characterized by OC-like spectra, but modified
by space weathering to the present-day darker and redder spectra. Labo-
ratory experiments (Sasaki et al., 2001; Brunetto et al., 2006)) and observa-
tions conducted by the NEAR and Hayabusa space missions (Clark et al.,
2002; Ishiguro et al., 2007) ) supported this theory.
A large number of Q-type asteroids have been detected in the near-Earth
region (Binzel et al., 2001; Dandy et al., 2003; Stuart and Binzel, 2004; DeMeo and Carry,
2013), the ratio of Q/S in NEAs is 0.5 to 2. If Q-type NEAs were produced
by collisions, we should also detect Q-type MBAs because the collisional rate
2
in main-belt is higher than that in near-Earth region. While, Q-type as-
teroids were missing in the main-belt in the earlier studies (Bus and Binzel,
2002a,b; Lazzaro et al., 2004), more recent observations have detected sev-
eral Q-type MBAs in the extremely young asteroid family "Datura dynam-
ical cluster" (Moth´e-Diniz and Nesvorn´y, 2008) and the older Koronis fam-
ily (Rivkin et al., 2011; Thomas et al., 2011). Carvano et al. (2010) classed
3296 of 62576 asteroids as Q-type-like objects in SDSS Moving Object Cata-
log (SDSSMOC4). Polishook et al. (2014) also detect two Q-type asteroids,
(19289) 1996 HY12 and (54827) 2001 NQ8, in the unbound asteroid pairs.
These results show that Q-type taxonomy is not limited to the NEA popula-
tion. However, the abundance of Q-type asteroids in the main-belt is still low
comparing with that in the near-Earth region. A simple explanation for this
low ratio of Q/S in MBAs is that the survey of small MBAs is incomplete,
many of Q-type MBAs might be discovered if the observations are able to
detect the sub-kilometer-sized MBAs.
The collisional formation model of Q-type NEAs (hereafter, "standard
model") could be challenged by the rapid process of space weathering effect
with solar wind implantation (Hapke, 2001; Vernazza et al., 2009), which
timescale could be as short as ∼ 104 to 106 years. Two arguments have been
presented. First, kilometer-sized or large asteroids with collisional lifetimes
exceeding 108 years (Bottke et al., 1993, 1994) should not display Q-type
spectra under long-term space weathering effect. However, the existence
of kilometer-sized Q-type NEAs, such as (1862) Apollo (Stuart and Binzel,
2004), contradicts the prediction of the "standard model". Second, the high
collisional rate in the main-belt region can produce asteroids with fresh
surfaces more efficiently. On the other hand, the time scale of transport
processes, such as the Yarkovsky effect and small resonances, that insert
collisional fragments into the planet-crossing space also exceeds 106 years
(Rabinowitz, 1997; Morbidelli and Vokrouhlick´y, 2003; Migliorini et al., 1998;
Bottke et al., 2002; Binzel et al., 2004). Therefore, Q-type NEAs should not
be present if they were primarily transferred from the main-belt with a Q-
type spectra. This contradicts the observations made in near-Earth space.
An alternative scenario involves a possibility that the surfaces of Q-type
NEAs have been reset during planetary encounters, from which the surface
materials were removed (Nesvorn´y et al., 2005) or re-arranged (Binzel et al.,
2010) by tidal effect. This hypothesis has been tested by several theoretical
and observational studies. For example, Marchi et al. (2006) determined
that the spectral slope of Q-type asteroids is correlated with planet-crossing
3
frequency. Binzel et al. (2010) and Nesvorn´y et al. (2010) suggested that the
Q-type NEAs have experienced encounters with the Earth, Venus and then
the tidal forces from these terrestrial planets could refresh the asteroidal
surfaces. DeMeo et al. (2014) proposed the possibility that this mechanism
might also be valid for Mars. As a corollary, the planetary encounter models
predicts that Q-type asteroids are rare among MBAs because of the low
planetary encounter rate in the main-belt.
Nevertheless, the timescale of space weathering on asteroid surface is
still in debate. Willman and Jedicke (2011) studied 95 asteroids for which
span a size and age range of about 1-20 km and 100-3000 Myr, respectively,
and measured a space weathering time of 2 × 109 years. This is much longer
than the result of fast space weathering (Hapke, 2001; Vernazza et al., 2009).
Polishook et al. (2014) also detected a Q-type asteroid in main-belt with age
∼ 106 years indicating that the space weathering timescale should be no less
than ∼ 106 years.
The other mechanisms to create Q-type asteroids are correlated with fast
rotation and YORP effect: 1. rotational-fission results in the exposure of
material from the covered surface of parent asteroid (Polishook et al., 2014),
and 2. rotational re-arrangement of asteroid surface material via landslips
(Scheeres, 2015; Walsh et al., 2012) and partial removal of weathered re-
golith. These two mechanisms are able to uncover non-weathered materials
and display the fresh Q-type spectra. Since the rotation of the smaller as-
troids are easier to be accelerated by YORP effect, we expect to detect more
small size Q-type MBAs if rotational effects are the dominant mechanism of
the Q-type asteroid formation.
Delbo et al. (2014) reported recently that thermal degradation induced
by diurnal temperature variation is able to break up rocks on the asteroid
surface rapidly into new regolith layer. They also suggested that asteroids
with large diurnal temperature difference (i.e., NEAs) can be cover by fresh
regolith characterized by the Q-type spectra. This scenario predicts that
more Q-type asteroids should be detected in near Earth space than in main-
belt because of the larger diurnal temperature variation of NEAs. Note
that the regolith formation by thermal fragmentation does not depend on
asteroid size; it may also imply that there is no color-size relation in the
NEA population.
From the discussion above, the multicolor observation of kilometer to sub-
kilometer diameter MBAs becomes critical to understand the space weath-
ering on S-complex asteroid surface and the formation of Q-type asteroids.
4
We should detect a comparable or even higher fraction of Q-type asteroids in
the main-belt than the near-Earth region, if the space weathering timescale
is > 108 years, and the collisional "standard model" dominate the forma-
tion of Q-type asteroids. By contrast, if the Q-type asteroid fraction in the
main-belt is very low, the Q-type NEAs must form in-situ and other mecha-
nism like the planetary encounter models, rotational-fission/re-arrangement
or thermal degradation should be responsible of the formation of Q-type
NEAs.
2. Observations and Data Reduction
To find sub-kilometer asteroids in the main-belt, we used the data taken
by Subaru telescope with Suprime-Cam, which is a prime focus camera with
a wide field of view (34' x 27') that consists of 10 CCD chips (Miyazaki et al.,
2002). The observational dates were August 9 and 10, 2004 (UT).
Three fields were surveyed each night for approximately 3.5 hours from
the midnight of Hawaii. The seeing size was 0.54-0.70 arcsecond on the
first night and 0.77-1.06 arcsecond on the second night. The observational
fields were near opposition and close to the ecliptic plane. The center of the
coordinates of each observed field is listed in Table 1.
The images were obtained using four broadband filters: B, V , R and I.
The exposure times were 120 sec for the B-, V - and R-bands and 180 sec for
the I-band. The observations followed the color sequence RR-BB-V V -RR-
II-RR for each fields. The time interval of the first R-band set and the second
R-band set was approximately 80 min. The interval between the second R-
band set and the third R-band set was approximately 60 min. We used
this three sets of R-band observations to interpolate the R magnitude in the
epoch of B, V and I-band observations to avoid possible color uncertainty
due to the asteroid rotational effect. Detailed description of photometric
calibration can be found in Sections 2.2 and 2.3 in detail.
5
Table 1 Coordinates of the surveyed fields
Aug. 09, 2004
Field ID
F1-1
F2-1
F3-1
Aug. 10, 2004
RA
21:18:24
21:18:24
21:18:24
DEC
λ
-15:11:00 317.306
-15:41:00 317.154
-16:11:00 317.003
β
0.488
0.011
-0.465
Field ID
RA
DEC
λ
F1-2
F2-2
F3-2
21:22:12
21:22:12
21:22:12
-14:54:00 318.266
-15:24:00 318.112
-15:54:00 317.959
β
0.478
0.002
-0.474
2.1. Detection of Moving Objects
To detect moving objects in relatively crowded fields, we first stacked all
12 exposures in each field to obtain deep images. We then used these deep
images as the source images to generate the reference stationary catalogs and
remove all stationary sources in every exposure.
After removing the stationary sources, we used the KDTree-based near-
est neighborhood search method to identify the detection pairs in every two
consecutive exposures. The pairs detected in the first set of consecutive expo-
sures were used to determine the main vectors for predicting the possible lo-
cations in the other five exposure pairs. We then searched for the correspond-
ing pairs at the predicted locations of the other set of consecutive exposure.
Once the entire set of six pairs (i.e., 12 detections in total) was identified,
the complete set was passed to the Orbf it code (Bernstein and Khushalani,
2000) to ensure that the orbital solution are reasonable; the semi-major axis
was between 2 AU and 5 AU, and the fitting residual was smaller than 0.5".
Under these stringent conditions, all moving objects detected are real and
complete color measurements were performed for all of them. The asteroid
detection list is summarized in Table 2.
6
RA (◦) DEC (◦) Epoch (MJD) a (AU)a i (◦)a HV
B
Berr
V
Verr
R
Rerr
I
Ierr
Ab Ab
err
Comment
a (AU)c
e c
i (◦)c
Table 2. Asteroid list
7
319.48272 -15.40419 53226.529043
319.39225 -15.22197 53226.529043
319.69678 -15.20883 53226.529043
319.46102 -15.20154 53226.529043
319.46236 -15.19466 53226.529043
319.58996 -15.18535 53226.529043
319.57579 -15.16933 53226.529043
319.38260 -15.16018 53226.529043
319.34533 -15.16815 53226.529043
319.63049 -15.14830 53226.529043
319.86432 -15.11508 53226.529043
319.37775 -15.09740 53226.529043
319.82910 -15.09154 53226.529043
319.58372 -15.09950 53226.529043
319.81048 -15.08243 53226.529043
319.72990 -15.05428 53226.529043
319.31365 -15.03409 53226.529043
319.86975 -15.03411 53226.529043
319.32668 -15.01941 53226.529043
319.70512 -14.98392 53226.529043
319.76720 -15.30467 53226.529043
319.88038 -15.29547 53226.529043
319.74818 -15.28862 53226.529043
319.30847 -15.28169 53226.529043
319.86194 -15.28194 53226.529043
319.71990 -15.27244 53226.529043
319.35283 -15.27298 53226.529043
319.83410 -15.23152 53226.529043
319.62572 -15.22084 53226.529043
319.69906 -15.22543 53226.529043
320.70343 -15.12162 53227.533671
320.65551 -15.05483 53227.533671
320.68576 -15.01347 53227.533671
2.84
2.90
3.21
3.09
2.25
2.55
2.84
3.05
2.56
2.92
2.53
2.85
3.46
2.61
2.99
2.91
2.71
2.91
2.92
2.29
3.28
2.41
2.68
3.16
2.97
3.03
2.72
3.01
2.91
3.06
2.54
2.99
2.88
18.103 22.561 0.014 21.690 0.002 21.277 0.011 20.926 0.008 0.088 0.011
2.31
19.612 24.146 0.021 23.320 0.002 22.981 0.017 22.584 0.001 0.004 0.106
4.69
0.03
18.980 23.970 0.025 23.231 0.037 22.980 0.038 22.661 0.001 -0.119 0.053
16.65 19.088 23.911 0.008 23.141 0.003 22.749 0.017 22.246 0.016 0.002 0.034
18.683 21.583 0.030 20.930 0.012 20.607 0.014 20.563 0.027 -0.130 0.024
5.50
19.353 23.149 0.001 22.336 0.026 21.922 0.013 21.560 0.012 0.048 0.048
8.45
9.71
18.900 23.027 0.012 22.493 0.006 21.859 0.012 21.519 0.023 0.006 0.020
0.16
19.211 23.882 0.087 23.200 0.015 23.056 0.032 22.667 0.020 -0.236 0.140
23.00 21.014 25.138 0.016 24.025 0.023 23.339 0.240 22.887 0.050 0.452 0.351
1.07
20.022 24.513 0.255 23.774 0.046 23.602 0.048 23.087 0.125 -0.176 0.245
11.95 15.351 19.177 0.002 18.279 0.003 17.823 0.003 17.423 0.009 0.138 0.003
2.08
19.851 24.105 0.030 23.461 0.012 23.022 0.064 22.618 0.041 -0.054 0.050
0.68
18.449 23.810 0.041 23.106 0.028 22.772 0.037 22.467 0.009 -0.086 0.044
19.12 18.133 22.075 0.010 21.258 0.003 20.762 0.013 20.367 0.004 0.109 0.008
13.25 19.227 23.837 0.042 23.096 0.028 22.738 0.049 22.321 0.026 -0.042 0.040
19.240 23.795 0.052 22.967 0.004 22.727 0.051 22.365 0.019 -0.065 0.104
0.46
19.777 23.905 0.011 23.108 0.006 22.726 0.015 22.375 0.057 0.014 0.045
7.15
4.80
15.174 19.783 0.008 18.904 0.000 18.410 0.003 17.941 0.002 0.151 0.008
17.177 21.680 0.006 20.918 0.002 20.463 0.007 20.018 0.020 0.041 0.006
1.91
5.76
19.579 22.853 0.033 21.928 0.013 21.657 0.003 21.035 0.024 0.025 0.031
10.14 17.388 22.444 0.001 21.760 0.002 21.409 0.008 21.071 0.003 -0.089 0.027
19.033 22.554 0.020 21.687 0.012 21.228 0.008 20.831 0.008 0.118 0.019
2.13
4.36
15.391 19.505 0.001 18.654 0.003 18.189 0.001 17.780 0.020 0.111 0.003 (62523) 2000 SW24
10.69 18.476 23.424 0.028 22.650 0.021 22.377 0.022 21.998 0.011 -0.080 0.041
8.94
19.373 23.631 0.085 23.205 0.163 22.552 0.029 22.227 0.050 -0.057 0.208
14.48 20.092 25.335 0.035 24.039 0.074 23.739 0.039 23.551 0.052 0.308 0.075
20.254 24.188 0.004 23.599 0.053 23.245 0.020 22.988 0.002 -0.153 0.064
4.47
17.812 22.470 0.004 21.713 0.015 21.393 0.016 21.670 0.434 -0.058 0.025
9.89
1.49
13.441 17.950 0.005 17.165 0.011 16.909 0.039 16.808 0.028 -0.084 0.018
19.460 24.116 0.056 23.453 0.063 23.019 0.036 22.602 0.004 -0.045 0.067
0.96
19.229 23.109 0.009 22.184 0.009 21.719 0.019 21.398 0.002 0.163 0.035
0.60
16.549 21.252 0.001 20.414 0.004 19.955 0.004 19.528 0.005 0.097 0.004
1.51
3.55
18.366 22.788 0.006 22.040 0.021 21.655 0.020 21.333 0.004 -0.019 0.016
2.7
0.07 4.28
Table 2 (cont'd)
RA (◦) DEC (◦) Epoch (MJD) a (AU)a i (◦)a HV
B
Berr
V
Verr
R
Rerr
I
Ierr
Ab Ab
err
Comment
a (AU)c
e c
i (◦)c
8
320.59647 -15.00993 53227.533671
320.33657 -14.97422 53227.533671
320.66292 -14.89674 53227.533671
320.44262 -14.85440 53227.533671
320.58374 -14.84279 53227.533671
320.26807 -14.82940 53227.533671
320.54242 -14.80967 53227.533671
320.52186 -14.78980 53227.533671
319.51576 -15.68548 53226.534619
319.85629 -15.67499 53226.534619
319.75825 -15.67583 53226.534619
319.67486 -15.66787 53226.534619
319.67821 -15.66154 53226.534619
319.38817 -15.64510 53226.534619
319.50101 -15.63361 53226.534619
319.70150 -15.63029 53226.534619
319.72054 -15.60861 53226.534619
319.36405 -15.57624 53226.534619
319.30469 -15.51043 53226.534619
319.30965 -15.50791 53226.534619
319.36096 -15.50186 53226.534619
319.54940 -15.87029 53226.534619
319.35954 -15.87175 53226.534619
319.60794 -15.85584 53226.534619
319.84755 -15.86667 53226.534619
319.80565 -15.84336 53226.534619
319.74746 -15.84407 53226.534619
319.49048 -15.83221 53226.534619
319.61032 -15.82195 53226.534619
319.46590 -15.82367 53226.534619
319.70470 -15.80420 53226.534619
319.47202 -15.79704 53226.534619
319.59842 -15.79135 53226.534619
3.03
3.04
3.05
2.64
3.12
2.88
2.94
3.11
2.50
2.57
2.31
2.14
3.23
2.68
2.41
2.19
2.60
2.92
2.93
2.52
2.54
2.43
2.69
2.61
3.00
3.13
2.65
2.84
2.64
3.11
2.43
2.77
2.77
10.02 19.492 24.327 0.030 23.437 0.034 23.172 0.025 22.874 0.016 -0.003 0.047
17.160 22.045 0.004 21.123 0.012 20.678 0.007 20.248 0.002 0.146 0.015 (252718) 2002 CQ194
2.47
19.167 24.113 0.025 23.148 0.110 22.830 0.077 22.386 0.042 0.087 0.092
0.55
19.728 24.209 0.063 22.907 0.191 22.512 0.049 22.306 0.037 0.380 0.220
3.27
16.584 21.393 0.009 20.680 0.009 20.331 0.005 19.975 0.001 -0.069 0.016
0.43
16.242 20.761 0.006 19.912 0.001 19.504 0.005 19.183 0.003 0.069 0.015 Q-type candidate
2.04
19.503 23.711 0.160 23.289 0.049 22.955 0.029 22.512 0.023 -0.285 0.152
9.75
18.405 23.366 0.013 22.495 0.016 21.906 0.078 21.411 0.022 0.213 0.031
2.37
20.173 23.905 0.032 23.036 0.001 22.675 0.011 22.290 0.010 0.050 0.033
9.54
20.294 24.496 0.009 23.330 0.062 23.168 0.013 22.646 0.090 0.119 0.062
6.24
21.302 24.414 0.004 23.709 0.069 23.215 0.022 22.703 0.048 0.028 0.108
0.80
22.288 24.940 0.189 24.237 0.004 23.481 0.063 23.408 0.000 0.212 0.158
1.03
19.251 23.904 0.309 23.531 0.024 23.224 0.036 22.888 0.035 -0.339 0.221
2.79
20.630 24.276 0.162 23.901 0.102 23.324 0.039 23.055 0.008 -0.146 0.141
8.59
14.861 18.618 0.007 17.524 0.012 17.207 0.014 16.954 0.012 0.178 0.013
3.41
16.856 19.749 0.004 18.932 0.003 18.346 0.043 18.092 0.003 0.172 0.005 (204290) 2004 PV21
7.19
17.202 21.007 0.002 20.289 0.006 19.983 0.006 19.638 0.000 -0.095 0.005
3.43
18.437 22.750 0.010 22.180 0.111 21.635 0.008 21.247 0.008 -0.031 0.106
7.74
19.062 23.533 0.026 22.831 0.012 22.508 0.013 22.208 0.022 -0.095 0.063
1.91
3.71
21.024 24.562 0.079 23.941 0.050 23.454 0.096 23.684 0.008 -0.037 0.143
17.38 19.236 22.934 0.044 22.198 0.017 21.540 0.014 21.088 0.000 0.166 0.040
19.608 23.105 0.043 22.312 0.003 21.859 0.025 21.406 0.009 0.062 0.034
6.49
0.02
20.663 24.784 0.009 23.956 0.019 23.692 0.048 23.171 0.112 -0.048 0.022
14.67 20.238 24.104 0.094 23.357 0.063 22.864 0.014 22.486 0.017 0.057 0.101
13.01 18.785 23.621 0.263 22.670 0.008 22.254 0.023 21.901 0.009 0.147 0.187
15.163 20.077 0.000 19.285 0.011 18.832 0.007 18.504 0.002 0.061 0.012
8.43
19.940 23.945 0.002 23.142 0.029 22.692 0.020 22.297 0.038 0.066 0.027
0.94
17.144 21.536 0.010 20.732 0.007 20.299 0.004 19.942 0.002 0.055 0.009
2.87
20.579 24.493 0.038 23.763 0.031 23.312 0.033 22.931 0.013 0.015 0.060
2.77
1.35
17.336 21.778 0.003 21.417 0.001 20.367 0.399 20.501 0.008 0.177 0.009
20.853 24.404 0.020 23.553 0.026 23.061 0.040 22.749 0.013 0.130 0.029
6.59
20.608 24.561 0.061 24.058 0.024 23.762 0.068 23.680 0.044 -0.255 0.061
5.62
3.68
19.621 23.884 0.015 23.078 0.008 22.683 0.032 22.294 0.047 0.030 0.048
3.01
0.08 3.47
2.17
0.1
6.36
RA (◦) DEC (◦) Epoch (MJD) a (AU)a i (◦)a HV
B
Berr
V
Verr
R
Rerr
I
Ierr
Ab Ab
err
Comment
a (AU)c
e c
i (◦)c
Table 2 (cont'd)
9
319.61758 -15.79291 53226.534619
319.45546 -15.78328 53226.534619
319.75107 -15.77375 53226.534619
319.71019 -15.75217 53226.534619
319.39739 -15.74895 53226.534619
319.43563 -15.73506 53226.534619
319.35384 -15.72188 53226.534619
320.42609 -15.57531 53227.539246
320.79772 -15.55019 53227.539246
320.71232 -15.53839 53227.539246
320.82430 -15.53347 53227.539246
320.32744 -15.53627 53227.539246
320.33516 -15.52610 53227.539246
320.67588 -15.49924 53227.539246
320.41969 -15.47944 53227.539246
320.58908 -15.46737 53227.539246
320.28343 -15.45079 53227.539246
320.80845 -15.38456 53227.539246
320.49664 -15.38502 53227.539246
320.63919 -15.37706 53227.539246
320.63703 -15.38200 53227.539246
320.54423 -15.37073 53227.539246
320.32559 -15.33817 53227.539246
320.76977 -15.30567 53227.539246
320.70791 -15.28156 53227.539246
320.45096 -15.27708 53227.539246
320.64892 -15.22799 53227.539246
320.63783 -15.20756 53227.539246
319.35421 -16.19792 53226.540177
319.60419 -16.39941 53226.540177
319.34255 -16.38858 53226.540177
319.62069 -16.37934 53226.540177
319.33731 -16.17518 53226.540177
2.23
2.49
2.71
2.47
2.77
2.64
2.80
2.22
2.72
3.04
2.54
2.28
2.38
3.16
2.37
3.30
2.72
3.00
2.89
3.30
2.57
2.68
3.05
3.02
2.34
2.24
2.45
2.72
3.11
2.42
3.78
2.82
2.84
20.205 22.574 0.211 22.403 0.213 21.716 0.010 21.393 0.017 -0.213 0.232
4.60
17.988 21.554 0.000 20.831 0.004 20.460 0.005 20.094 0.001 -0.047 0.005
4.66
20.252 24.560 0.019 23.586 0.039 23.325 0.037 22.823 0.042 0.053 0.048
3.71
19.898 23.581 0.031 22.691 0.014 22.219 0.011 21.775 0.044 0.143 0.042
1.68
18.324 22.572 0.030 21.779 0.007 21.291 0.007 20.898 0.001 0.086 0.022
6.53
20.248 24.317 0.005 23.435 0.014 23.026 0.017 22.388 0.156 0.093 0.037
1.59
18.519 22.810 0.017 22.026 0.005 21.627 0.028 21.333 0.010 0.017 0.013
5.80
15.012 17.950 0.005 17.171 0.005 16.604 0.007 16.435 0.037 0.132 0.008
3.42
18.238 22.254 0.019 21.593 0.014 21.240 0.014 20.890 0.024 -0.103 0.024
4.30
18.195 22.782 0.007 22.164 0.023 21.677 0.063 21.371 0.003 -0.039 0.029
2.51
19.103 22.670 0.042 22.071 0.034 21.274 0.062 20.832 0.022 0.167 0.085
3.91
17.616 20.842 0.008 19.942 0.017 19.459 0.004 19.066 0.003 0.158 0.014
3.97
3.91
18.111 21.565 0.002 20.702 0.018 20.258 0.005 19.823 0.014 0.104 0.013
10.05 18.341 23.253 0.030 22.519 0.004 22.227 0.004 21.915 0.009 -0.094 0.034
20.047 23.933 0.117 22.608 0.009 22.424 0.060 22.051 0.018 0.247 0.091
7.19
15.615 20.736 0.013 20.020 0.001 19.653 0.023 19.338 0.013 -0.054 0.009
4.38
17.285 21.499 0.001 20.642 0.010 20.278 0.007 19.797 0.055 0.043 0.010
5.32
15.556 20.261 0.008 19.456 0.002 19.141 0.005 18.844 0.005 -0.028 0.006
7.79
2.26
19.593 24.077 0.132 23.271 0.006 22.923 0.216 22.194 0.003 -0.004 0.099
14.74 17.682 23.284 0.057 22.088 0.092 22.118 0.010 21.781 0.005 0.004 0.090
10.28 18.101 22.082 0.130 21.137 0.001 20.780 0.011 20.310 0.012 0.100 0.093
19.424 23.435 0.044 22.698 0.013 22.313 0.014 21.963 0.012 -0.027 0.083
6.37
17.527 22.273 0.004 21.516 0.019 21.192 0.005 20.675 0.000 -0.055 0.014
1.41
0.90
17.848 22.515 0.017 21.767 0.024 21.427 0.007 21.112 0.009 -0.051 0.024
18.318 21.534 0.002 20.805 0.001 20.412 0.006 20.038 0.007 -0.027 0.005
1.87
17.172 20.377 0.005 19.389 0.008 18.891 0.010 18.554 0.021 0.231 0.009 (151733) 2003 BA88
1.83
19.498 23.051 0.010 22.241 0.035 21.872 0.023 21.580 0.041 0.014 0.030
2.77
18.807 22.916 0.029 22.153 0.025 21.774 0.014 21.434 0.021 -0.012 0.037
0.08
18.054 22.743 0.020 22.135 0.010 21.703 0.009 21.590 0.001 -0.084 0.024
0.65
4.00
20.306 23.827 0.051 22.981 0.184 22.668 0.397 22.918 0.070 0.000 0.173
13.11 17.136 23.051 0.021 22.249 0.008 21.799 0.021 21.476 0.024 0.065 0.021 Q-type candidate
9.67
2.81
18.962 23.206 0.054 22.507 0.026 22.084 0.015 21.272 0.299 -0.027 0.047
20.090 24.511 0.038 23.689 0.048 23.318 0.080 23.047 0.076 0.024 0.098
(6919) Tomonaga
(2014) AR8
2.26
0.1
5.16
2.4
2.32
0.14 1.72
0.15 2.64
RA (◦) DEC (◦) Epoch (MJD) a (AU)a i (◦)a HV
B
Berr
V
Verr
R
Rerr
I
Ierr
Ab Ab
err Comment a (AU)c
e c
i (◦)c
Table 2 (cont'd)
1
0
319.57251 -16.37746 53226.540177
319.86114 -16.36801 53226.540177
319.64018 -16.13018 53226.540177
319.54882 -16.12899 53226.540177
319.32721 -16.11899 53226.540177
319.60643 -16.12420 53226.540177
319.82640 -16.10238 53226.540177
319.51692 -16.09192 53226.540177
319.83680 -16.07527 53226.540177
319.76208 -16.06889 53226.540177
319.72250 -16.06501 53226.540177
319.36741 -16.05471 53226.540177
319.47395 -16.04208 53226.540177
319.85396 -16.35700 53226.540177
319.72944 -16.02779 53226.540177
319.73087 -16.35587 53226.540177
319.76130 -16.02214 53226.540177
319.58537 -16.01515 53226.540177
319.61491 -16.00706 53226.540177
319.43173 -16.00280 53226.540177
319.72263 -16.35138 53226.540177
319.62979 -15.96643 53226.540177
319.51548 -15.97303 53226.540177
319.69070 -15.97569 53226.540177
319.45615 -16.33313 53226.540177
319.69838 -16.32601 53226.540177
319.69109 -16.28867 53226.540177
319.85657 -16.27948 53226.540177
319.85490 -16.26903 53226.540177
319.56316 -16.23162 53226.540177
319.42440 -16.20991 53226.540177
320.79963 -16.11936 53227.544813
320.56040 -16.08520 53227.544813
2.62
2.35
2.98
2.77
2.95
2.59
2.59
3.02
2.35
2.96
2.67
3.06
3.07
2.91
2.52
2.75
2.68
2.65
2.75
3.02
2.98
2.81
2.39
3.02
2.99
2.51
2.63
3.12
2.68
3.22
2.96
2.99
2.04
18.422 22.469 0.004 21.555 0.006 21.091 0.008 20.694 0.008 0.155 0.015
2.94
1.14
20.808 24.117 0.028 23.307 0.025 22.627 0.025 22.147 0.018 0.234 0.047
12.47 18.580 23.257 0.085 22.428 0.008 22.049 0.007 21.657 0.002 0.035 0.061
6.70
19.475 23.585 0.022 22.919 0.019 22.515 0.041 22.737 0.081 -0.064 0.036
15.29 16.878 21.484 0.003 20.680 0.007 20.161 0.011 19.713 0.002 0.115 0.014
18.688 22.621 0.009 21.759 0.003 21.339 0.008 20.964 0.009 0.087 0.014
3.05
20.109 23.909 0.049 23.185 0.051 22.704 0.034 22.295 0.010 0.032 0.064
3.40
7.10
18.453 23.231 0.004 22.387 0.012 21.858 0.017 21.456 0.011 0.150 0.025
2.16
19.252 22.467 0.011 21.753 0.005 21.402 0.014 21.058 0.002 -0.067 0.011
11.26 18.247 22.939 0.010 22.072 0.096 21.766 0.015 21.404 0.007 0.009 0.103
18.203 22.303 0.007 21.451 0.013 21.008 0.007 20.626 0.008 0.096 0.019
7.38
18.246 23.006 0.009 22.250 0.034 21.893 0.009 21.517 0.008 -0.033 0.067
7.98
19.673 24.456 0.005 23.684 0.012 23.343 0.016 22.974 0.008 -0.033 0.055
8.63
1.61
18.683 23.172 0.046 22.411 0.039 22.125 0.032 21.742 0.005 -0.080 0.046
19.079 22.897 0.021 22.000 0.007 21.660 0.026 21.253 0.010 0.055 0.024
6.36
16.955 21.061 0.009 20.376 0.012 19.827 0.005 19.377 0.020 0.052 0.014
1.16
20.524 24.626 0.030 23.783 0.031 23.335 0.025 23.170 0.004 0.093 0.108
1.49
20.121 24.249 0.022 23.332 0.076 22.790 0.057 22.347 0.099 0.212 0.083
1.79
4.22
19.611 23.959 0.006 23.030 0.025 22.592 0.022 22.195 0.011 0.147 0.021
13.083 17.621 0.007 17.015 0.005 16.939 0.030 16.792 0.008 -0.338 0.007
0.46
19.868 24.295 0.106 23.729 0.042 23.196 0.029 22.892 0.087 -0.043 0.107
5.89
17.485 21.699 0.010 21.014 0.007 20.651 0.617 20.661 0.320 -0.079 0.014
2.03
17.151 20.531 0.005 19.769 0.002 19.490 0.010 19.097 0.002 -0.083 0.004
6.87
9.71
14.657 19.295 0.004 18.576 0.005 18.222 0.006 17.868 0.002 -0.061 0.005
12.99 18.454 23.144 0.004 22.319 0.009 21.933 0.025 21.498 0.031 0.036 0.015
20.177 24.522 0.018 23.069 0.406 23.014 0.050 22.693 0.037 0.247 0.303
0.48
20.354 23.898 0.198 23.519 0.050 23.163 0.021 22.787 0.005 -0.300 0.147
6.04
19.148 23.844 0.026 23.243 0.003 22.847 0.027 22.483 0.043 -0.115 0.024
0.50
20.756 24.994 0.056 24.023 0.117 23.626 0.089 23.131 0.099 0.147 0.167
2.55
4.24
15.270 20.221 0.006 19.547 0.020 19.088 0.012 18.382 0.013 -0.019 0.021
20.773 25.443 0.034 24.592 0.056 24.029 0.040 23.412 0.061 0.179 0.079
0.96
18.126 22.710 0.008 21.997 0.018 21.638 0.034 21.361 0.009 -0.062 0.041
1.87
5.53
21.185 23.541 0.010 22.824 0.023 22.194 0.017 21.742 0.022 0.132 0.024
2.2. Flux Calibration
The Suprime-Cam image data were calibrated chip by chip in every expo-
sure by identifying Pan-STARRS 1 (hereafter PS1) catalogue stars in the ob-
served field; the data were calibrated using "uber-calibration" (Magnier et al.,
2013). Uber-calibration is an algorithm used to photometrically calibrate
wide-field optical imaging surveys, and it was first applied to the Sloan Digital
Sky Survey imaging data. It can be used to simultaneously solve for the cal-
ibration parameters and relative stellar fluxes through the use of overlapping
observations (Padmanabhan et al., 2008). Those uber-calibrated catalogues
have a relative precision (compared with the SDSS) of < 10 mmag in gP 1,
rP 1, and iP 1, and approximately 10 mmag in zP 1 and yP 1 (Schlafly et al.,
2012). Since we used the B, V , R, I filters in our survey, we transferred
those PS1 catalogue stars from the PS1 gP 1rP 1iP 1 photometry system to our
Johnson-Cousins BV RI system by using the transformation equation and
parameters obtained by Tonry et al. (2012). The transformation error was
approximately 0.03 in the B-band and 0.01 to 0.02 in the V -, R-, I-bands.
2.3. Trail Fitting
Two problems occur when using the traditional photometry methods for
asteroid flux measurements: First, the flux would be contaminated by nearby
stars if aperture photometry is used to measure the asteroid flux within
crowded fields. Second, PSF fitting fails to yield accurate results if the as-
teroid image is not point-source-like object under long exposure. Therefore,
we applied the trail fitting technique to measure the asteroid flux.
The trail function chosen is an axisymmetric Gaussian PSF-convolution
trail function given in equation (3) of Veres et al. (2012):
g[x′, y′] = b +
F
L
1
2σ (2π)1/2 exp(cid:20)−
y′2
2σ2(cid:21)(cid:26)erf (cid:20) x′ + 1
2 L
σ√2 (cid:21) − erf (cid:20)x′ − 1
σ√2 (cid:21)(cid:27) ,
2L
where
and
erf [z] =
2
π Z z
0
exp(cid:2)t2(cid:3)dt
x′ = (x − xc) cos θ + (y − yc) sin θ,
y′ = −(x − xc) sin θ + (y − yc) cos θ,
11
(1)
(2)
(3)
(4)
RA (◦) DEC (◦) Epoch (MJD) a (AU)a i (◦)a HV
B
Berr
V
Verr
R
Rerr
I
Ierr
Ab Ab
err Comment a (AU)c
e c
i (◦)c
Table 2 (cont'd)
320.64800 -16.03000 53227.544813
320.27458 -16.02996 53227.544813
320.40652 -16.01986 53227.544813
320.31635 -16.00107 53227.544813
320.82385 -15.98608 53227.544813
320.71048 -15.93982 53227.544813
320.65166 -15.91541 53227.544813
320.62216 -15.91270 53227.544813
320.39658 -15.87950 53227.544813
320.41451 -15.86789 53227.544813
320.39999 -15.85829 53227.544813
320.70450 -15.85786 53227.544813
320.66966 -15.78510 53227.544813
320.65956 -15.74312 53227.544813
320.51081 -15.74918 53227.544813
320.65568 -15.73321 53227.544813
320.32995 -15.71095 53227.544813
320.30033 -15.68513 53227.544813
2.87
3.05
2.29
2.25
2.44
2.64
2.33
3.13
2.56
2.72
3.10
2.64
3.22
2.91
2.69
2.51
3.08
3.15
17.42 17.010 21.444 0.009 20.651 0.006 20.265 0.005 19.901 0.006 0.014 0.009
7.99
17.404 22.289 0.046 21.376 0.014 20.940 0.011 20.564 0.024 0.134 0.035
12.74 20.260 23.424 0.047 22.600 0.020 22.209 0.028 21.855 0.002 0.039 0.047
19.489 22.591 0.019 21.726 0.000 21.300 0.008 20.938 0.004 0.093 0.026
0.53
18.920 22.623 0.021 21.652 0.010 21.032 0.009 20.621 0.007 0.305 0.021
3.24
19.094 23.258 0.031 22.283 0.004 21.818 0.010 21.453 0.019 0.198 0.028
4.53
20.093 23.306 0.004 22.550 0.007 22.065 0.029 21.109 0.511 0.057 0.006
7.10
16.622 21.451 0.007 20.738 0.007 20.321 0.004 19.957 0.001 -0.021 0.008
3.21
2.40
19.464 23.241 0.008 22.462 0.062 22.075 0.010 21.649 0.004 0.004 0.062
18.944 22.866 0.126 22.304 0.055 21.900 0.047 21.499 0.038 -0.137 0.119
3.93
18.993 23.920 0.041 23.057 0.044 22.648 0.020 22.590 0.149 0.079 0.048
0.75
19.097 23.084 0.035 22.283 0.010 21.895 0.013 21.556 0.000 0.021 0.040
1.98
1.12
15.826 20.814 0.018 20.095 0.006 19.707 0.004 19.393 0.004 -0.037 0.014
24.42 17.230 21.705 0.018 20.951 0.000 20.661 0.006 20.222 0.001 -0.082 0.015
19.273 23.438 0.014 22.559 0.000 22.276 0.032 21.796 0.000 0.002 0.018
6.83
3.04
19.194 23.230 0.011 22.099 0.010 21.434 0.127 21.270 0.001 0.450 0.027
12.77 18.356 23.137 0.015 22.386 0.037 22.007 0.007 21.630 0.004 -0.021 0.032
10.08 18.531 23.528 0.019 22.691 0.000 22.247 0.010 21.592 0.191 0.086 0.030
1
2
Note. -- a Estimated by using Equation 6, 7,8 and 9 under assuming e=0.
Note. -- b See Equation 13.
Note. -- c Orbital elements of known asteroids are provided by JPL Small-Body Database.
where b is the background level, L is the length of the trail, F is the total
integrated flux in the trail, σ is the standard deviation of the PSF Gaussian,
xc and yc are the coordinates of the centroid of the trial. θ is the angle
between the long axis of the trail and the x axis.
We used the Levenberg-Marquardt least-squares fitting technique to min-
imize the variance between the image and trail function. Figure 1 shows an
example of an asteroid trail and its fitting residue. Clearly, the asteroid trail
can be completely subtracted by using the trail function. The photometric
error of the asteroid flux was estimated from the flux difference between two
consecutive exposures of the asteroid; we considered the flux difference to
be closer to the actual uncertainty compared with the estimation derived
from the fitting result. We expect that the SNR of trail fitting photometry
decreases with (2.35σ/(2.35σ + L))0.5. Thus the faster mover have larger
photometry uncertainty. Fortunately, there is no systematic effect related to
the trail length (See Veres et al. (2012) for more detail), and the photometry
results should be independent with the semi-major axis of asteroids.
2.4. Detection Efficiency and De-biasing
For testing the detection efficiency of moving objects in the observations,
we planted 500 non-trailing artificial stars with a brightness between 22.5
to 25.5 magnitude in each CCD chip. The artificial objects were generated
by IRAF task DAOP HOT by modeling the stars in each CCD chip in
each exposure. The limiting magnitude also depends on the trailing rate of
asteroids. Assuming that the moving rate of asteroids is about 0.01"/s in
average. We expect that the trail length is around 1.2" in the 120s exposure.
For the average 0.7" seeing size of our data, the SNR of the asteroid trail is
roughly 60% of the non-trailing source with the same brightness. Therefore,
we increased the detection threshold of artificial non-trailing stars to 5 sigma
to simulate the limiting magnitude of 3 sigma asteroid detections.
We counted the number of detected artificial objects with a function of
magnitude and plotted a diagram of magnitude vs.
fractional detection as
shown in Figure 2, and fitted a detection efficiency function, which was
a function with double hyperbolic tangents (Petit et al., 2006), to the test
result:
η(R) =
A
4
[1 − tanh(
R − Rc
∆1
)][1 − tanh(
R − Rc
∆2
)]
(5)
13
0
5
10
15
0
5
10
15
0
5
10
15
Figure 1 An example of an asteroid trail in our sample before (top) and after
(bottom) subtracting its fitting result (contour).
14
1.0
0.8
0.6
0.4
0.2
n
o
i
t
c
e
t
e
d
l
a
n
o
i
t
c
a
r
F
0.0
22.5
23.0
23.5
24.0
R-mag
24.5
25.0
25.5
Figure 2 An example of fractional detection as a function of the R-magnitude.
The best-fit detection efficiency function is illustrated.
Here, the fitted parameters A, Rc, ∆1 and ∆2 are the filling factor (or max-
imal efficiency), roll-over magnitude (50% of the maximal efficiency), and
widths of the two components, respectively. An exemplary result for the
R-band, Chip2, took on August 10, 2004 was the 50% detectability of ap-
proximately 23.7 magnitude, and the filling factor was approximately 83%
(see Figure 2).
Using this detection efficiency function, we assigned a weight for each
asteroid detection; the weight was the multiplicative inverse of the product of
the detection efficiency of every chip and exposure of the asteroid that passed.
The weight of the object with the highest detectability is 1, and fainter
objects, which have a low detection probability, have a higher weighting
value. Therefore, we eliminated the observational bias in the detection of
objects and tested the completeness of the survey.
3. Results
3.1. Absolute Magnitude and Completeness
We extracted a total of 150 asteroids with 12 detections from the six
fields and measured their apparent velocities. Assuming that their orbital
15
eccentricities were zero and they located around opposition during the obser-
vations, we estimated the semi-major axes and inclinations by using the fol-
lowing equations (Bowell and Lumme, 1979; Nakamura and Yoshida, 2002;
Yoshida and Nakamura, 2007):
1
2γ
a =
tan I =
(γ − 2kλ ±p κ ),
β
λ + k/(a − 1)
,
γ = λ2 + β2,
κ = γ2 − 4kλγ − 4k2β2,
(6)
(7)
(8)
(9)
where λ and β are the moving rate along the ecliptic longitude and latitude,
respectively; a and I are the semi-major axis and inclinations, respectively;
and k is the Gaussian gravitational constant. The semi-major axis and in-
clinations estimated using the aforementioned equations included errors of
approximately 0.1 AU and 5◦ because the actual eccentricities of the aster-
oids were not zero. We compared the orbital elements of known asteroids
in Table 2 with our estimated elements. The result shows that the estimate
elements are generally accurate; the difference between known and estimated
semi-major axes, inclinations are less than 0.1 AU, 2◦, respectively.
The absolute magnitude (HV ) of each asteroid at opposition can be esti-
mated using the following equation:
HV = m − 5log(∆r)
(10)
where m is the apparent magnitude of the asteroid and ∆ and r are geocentric
distance and heliocentric distance, respectively. For simplicity, r = a and
∆ = a − 1 because of the e = 0 and nearly opposition assumptions.
Figure 3 shows the absolute magnitude (HV ) distribution of our samples.
The black spikes are the raw HV distribution, and the gray boxes shows the
weighted result. Based on the results of Yoshida and Nakamura (2007), we
plotted two power laws (N (< HV ) ∝ 10αHV ) to the HV distribution, with
power law indices b of 1.29 (for the HV range from 17.8 to 20.2 mag) and 1.75
(14.6 to 17.4 mag). Here b = 5 α. This plot clearly indicates that our samples
have power law indices consist with the results of Yoshida and Nakamura
(2007) and satisfactory completeness of up to HV = 20, which corresponds to
asteroids with diameters smaller than 590 meter for the C-complex asteroids
16
weighted
unweighted
Figure 3 Weighted (Grey) and unweighted (black) cumulative absolute mag-
nitude (HV ) distributions of the asteroids sample. Solid and dashed lines
show the power laws distribution with power law indices of 1.29 (for the HV
range 17.8 to 20.2 mag) and 1.75 (14.6 to 17.4 mag), respectively.
(assuming an albedo of 0.05) and smaller than 270 meter for the S-complex
asteroids (assuming an albedo of 0.25).
3.2. Colors and Estimated Taxonomic Types
From our sample (150 asteroids), 75 asteroids with the HV range from 16
to 20 magnitude, the color errors smaller than 0.1, 0.2 < V − R < 0.6 and
0.6 < B − V < 1 was selected. These asteroids have satisfactory HV com-
pleteness and photometric accuracy which made them suitable for taxonomic
estimations. The color errors are propagated from the uncertainly of BVRI
photometry of asteroids in Table 2. Figure 4 shows the relative reflectances
in our asteroid samples overlap with the transmission curve of the Suprime-
Cam filter system. The C-complex asteroids generally have a flat reflectance
across the four filters, and the S-complex and V/Q/R-type asteroids exhibit
similar slopes in the B, V, R region but different reflectances near the I-band;
the V/Q/R-type asteroids demonstrated a stronger absorption feature at ∼
1 µm. Therefore, while B − V and V − R color are used to separate S-
complex and C-complex asteroids, the R − I color is needed to distinguish
S-type asteroids from V/Q/R-type-like asteroids in the following sense. The
17
e
c
n
a
t
c
e
l
f
e
R
1.2
1.1
1
.9
.8
.7
.6
B
V
Rc
Ic
0.4
0.5
0.6
0.7
Wavelenght (µm)
Q-type
S-type
C-type
0.8
Figure 4 Spectrophotometry of three types of asteroid spectra: S-complex,
C-complex and V/Q/R-type-like asteroids. The background shows the trans-
mission curves of the Suprime-Cam filter system.
S-type asteroids are those with R − I > (R − I)⊙ (0.332 ± 0.008, according
to Holmberg et al. (2006)); otherwise the V/Q/R-type-like asteroids.
We also separated our sample into large and small asteroids by HV =
18.5, which approximately corresponds to a diameter of 1 km for C-complex
asteroids. Figure 5 shows B−V vs. V −R color-color diagram for part of our
large asteroid sample (Hv < 18.5). The bimodular distribution corresponds
to the color difference between the C-complex and S-complex asteroids.
Moreover, we identified principal components (the uncorrelated variables)
for the large asteroid sample in B − V and V − R space by prcincipal compo-
nent analysis to separate S-complex asteroids and C-complex asteroids and
resulted two linear combinations of B − V and V − R colors:
P C1 ≡ 0.7071(B − V ) + 0.7071(V − R),
(11)
and
P C2 ≡ 0.7071(B − V ) − 0.7071(V − R)
(12)
The result shows that B − V and V − R axes become PC1 and PC2, respec-
tively, after rotating 45◦ counterclockwise.
Furthermore, there is a dip at PC1 = 0.82 in the histogram of PC1 and
can be the boundary of C and S-complex asteroids. Therefore, we define a
18
Hv<18.5
A
0.65
0.6
0.55
0.5
0.45
0.4
0.35
0.3
0.25
R
V
-
0.2
0.6
0.65
0.7
0.75
0.85
0.9
0.95
1
0.8
B-V
Figure 5 The B − V and V − R color-color diagram for our large asteroid
sample (HV < 18.5). The solid line shows the direction of new axis (A) of
the color-color distribution.
new axis 'A':
A ≡ 0.7071(B − V ) + 0.7071(V − R) − 0.82
(13)
Asteroids with A > 0 should belong to S-complex, otherwise are C-complex.
The new axis 'A' is shown as a solid line in Figure 5.
Figure 6 shows the color-color diagram of axis A vs. R − I. Each dot
represents an asteroid. A value of 0.332, which is corresponding to the solar
color, was subtracted from the R−I axis to separate S-complex and V/Q/R-
type. Two major points can be made based on Figure 6: (1) only two objects
can be certified as Q-type MBA candidates; (2) the small sample (HV > 18.5)
appeared to be more concentrated at A ≈ 0, though the C- and S-complexes
are clearly separated in large sample (Hv < 18.5).
3.3. Fraction of Each Estimated Taxonomic Type
To estimate the taxonomic type of asteroids, we considered A < 0 as
C-complex asteroids, A > 0 and R − I − 0.332 > 0 as S-complex and A > 0
and R − I − 0.332 < 0 as V/Q/R like asteroids. Four objects locate around
D/T type region were removed by visually selection. Figure 7 shows the
semi-major axis vs inclination distribution of the 75 asteroids with estimated
19
0.25
0.2
0.15
0.1
0.05
0
-0.05
-0.1
-0.15
X
X
X
G
G
C
C
F
F
B
B
Hv < 18.5
S
S
R
R
D
D
T
T
Q
Q
V
V
-0.2
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
A
D
Hv>18.5
X
S
C
C
B
Q
V
R
0.25
0.2
0.15
0.1
0.05
0
-0.05
-0.1
-0.15
2
3
3
.
0
-
I
-
R
2
3
3
.
0
-
I
-
R
-0.2
-0.2
-0.15
-0.1
-0.05
0
0.05
0.1
0.15
0.2
0.25
A
Figure 6 The R − I vs. A color-color diagrams of the large (upper) and
small (bottom) asteroid samples in the present Subaru data. The regions
corresponding to three types of asteroids are shown in the plots. The average
colors of B, C, D, F, G, S, T, V, Q, R and X-type asteroids from Dandy et al.
(2003) are also marked.
20
)
e
e
r
g
e
d
(
n
o
i
t
a
n
i
l
c
n
i
30
25
20
15
10
5
0
2
S-complex
C-complex
V/Q/R
T/D
2.5
3
3.5
4
semi-major axis (AU)
Figure 7 Semi-major axis (a) vs inclination (i) distribution of the 75 asteroids
with estimated taxonomic type. Gray dots show the a, i distribution of of
known asteroids.
taxonomic type. The distribution looks normal; there are more C-complex
than S-complex asteroids in outer main-belt. There is a possible V/Q/R-
type candidate located on 3.8 AU. It could because of wrong estimation in
semi-major axis.
To calculate the fraction of each estimated taxonomic type, we took the
uncertainly of boundaries of A and R − I axes, which are 0.02 from the bin-
size of A histogram and 0.008 from the uncertainty of (R− I)⊙, respectively,
and the uncertainly of asteroid color measurements into account. We use 2D
gaussian distribution with a covariance matrix equal to
(cid:18) A2
err
0
0
err (cid:19)
(R - I)2
to represent the probability distribution of each object in A, R − I space.
Therefore, the fraction of a object in a specific taxonomic type is its prob-
ability distribution multiply a Complementary Error Function, which is the
cumulative function of a gaussian distribution with the mean value equal to
boundary value and standard deviation equal to the uncertainty of boundary.
Figure 8 shows the fraction of asteroid types in our weighted and unweighted
sample.
21
n
o
i
t
c
a
r
f
0.6
0.5
0.4
0.3
0.2
0.1
0
Weighted
Unweighted
C
S
V/Q/R
Figure 8 Histograms of the fractions of estimated asteroid spectral types.
C and S are C-complex and S-complex asteroids, respectively, and V/Q/R
indicates V/Q/R-type-like asteroids.
Since our sample exhibits a HV range similar to that of the near-Earth
asteroid sample used by Dandy et al. (2003), it is worth to compare the
fraction of asteroid's type to that of NEA population. The Q/S ratio in
our observation is less than 0.05 in the main-belt region. By contrast, the
Q/S ratio in the near earth space is about 0.5 (Binzel et al., 2004) to 2
(Dandy et al., 2003), which is much higher than in the main-belt.
Finally, we tested the correlation between HV and R − I of S-complex
asteroids. As shown in Figure 9, unlike the result of Dandy et al. (2003) in
which the absorption band depth was correlated with S-complex NEA size,
we did not find any evidence that R− I color has significant correlation with
the size of S-complex MBA.
4. Discussion
There are only two possible Q-type asteroids candidates in our km to sub-
km sized MBA sample (see Figure 6). This fact indicates that the Q-type
asteroids are rare in the main-belt. The Q/S ratio is less than 0.05, which
is significantly lower than the value of NEAs (∼ 0.5 in Binzel et al. (2004)
and ∼ 2 in Dandy et al. (2003)). This result is also comparable with the
22
0.35
0.3
0.25
0.2
0.15
0.1
0.05
0
-0.05
-0.1
-0.15
2
3
3
.
0
-
I
-
R
-0.2
16
16.5
17
17.5
18.5
19
19.5
20
18
HV
Figure 9 Variation of R−I color (an indicator of absorption band depth) as a
function of the absolute magnitude, Hv, for the observed S-complex MBAs.
Q-type-like fraction in Carvano et al. (2010), which the main belt Q/S ratio
is 325/3784 ∼ 0.09 for the good classified asteroids in SDSSMOC4.
Since most of S-complex MBAs are weathered, we can estimate the upper
limit of space weathering timescale in main belt by using the collisional size-
age relation in Bottke et al. (2005) and Willman and Jedicke (2011):
Ts(yr) = 3.23 × 10(11.0−0.18Hv )
(14)
The Hv range of our MBA samples are between 16 to 20 magnitude, and
their corresponding collisional ages are between 8 × 107 to 4× 108 years. We
can safely conclude that the space weathering timescale in main belt should
be less than 108 years.
The lack of Q-type MBAs also suggests two facts: 1. Most of the Q-
type NEAs did not come from main-belt, they must form in-situ, and 2.
collision is not the main mechanism of the formation of Q-type NEAs due
to the collision rate being lower in the near Earth region. There must exist
other mechanisms to generate such large amount of Q-type NEAs, and these
mechanisms are more effective in the near Earth region than the main-belt.
The planetary encounter model (Binzel et al., 2010; Nesvorn´y et al., 2010)
advocating the recent resetting of S-type asteroid surfaces by the effects of
tidal stress is one of the possible mechanisms.
23
Another possible mechanism that could be responsible for the formation
of Q-type NEAs is YORP effect spin-up induced rotational-fission or sur-
face re-arrangement of asteroids. The acceleration rate of asteroid spin by
the YORP effect is inversely proportional to the square of semi-major axis
and more effective in the near Earth region than in the main-belt due to
the smaller heliocentric distance (Rubincam, 2000; Scheeres, 2007). Thus,
if rotational-fission mechanism or rotational re-arrangement is also several
times more effective to create Q-type NEAs than Q-type MBAs, it may be
able to explain why the Q/S ratio in NEAs is about 10 to 40 times larger
than Q/S ratio in MBAs.
The YORP spin-up can also explain the existence of main-belt Q-type
asteroids (see Polishook et al. (2014) for detail). It indicates the size-color
(or size-S/Q ratio) dependence of S-complex MBAs, because the smaller as-
teroid is easier to be accelerated to near the break-up limit resulting in the
Q-type-like color. However, such relation is not shown in our sample. There
are two possible effects that may cause the lack of size dependence:
1. The "secondary fission" of rotational-fission models provided by Jacobson and Scheeres
(2011) may be the more likely model to create main-belt Q-type asteroids.
This model implies the destruction of the secondary of pair asteroids by pri-
mary's tidal forces. For the km to sub-km sized asteroid pair, the gravity
may be too small to deform secondary and produce Q-type surface.
2. A size-dependent strength for asteroids in addition to gravity (Holsapple,
2007) can prevent the break-up of small asteroids. The existence of sub-km
sized super-fast rotator, such as (29075) 1950 DA (Rozitis et al., 2014) and
(335433) 2005 UW163 (Chang et al., 2014), might be the evidence of the
existence of this internal strength.
The other mechanism is thermal degradation of the rocks on asteroid
surface from Delbo et al. (2014). This process is strongly dependent on the
value of diurnal temperature difference, which is a function of perihelion
distance; it takes ∼ 105 yrs in near Earth region and ∼ 107 yrs in the main-
belt to break 90% of 3 centimeter diameter size rocks. If thermal degradation
dominates the formation of Q-type asteroids, space weathering must have
timescale < 107yrs to keep low Q/S ratio in the main-belt.
5. Summary
We surveyed kilometer- to sub-kilometer-sized asteroids in the main belt
by using the Subaru telescope. A total of 150 asteroids with BVRI colors
24
were detected and 75 of them exhibited satisfactory photometry accuracy.
The main results can be summarized as follows:
1. Q-type asteroids are rare in the main-belt; only two Q-type candidates
were detected in our sample, and the Q-type to S-type ratio is less than 0.05
in main-belt.
2. Unlike the size-color dependence of NEAs found by Dandy et al. (2003),
we did not found any evidence of that in MBA population.
3. The space weathering timescale in the main belt should be less than 108
years.
4. Re-arrangement of surface material of S-type asteroid by tidal stress dur-
ing planetary encounters and thermal degradation are possible mechanisms
of Q-type NEAs formation. YORP spin-up induced rotational-fission or sur-
face re-arrangement of asteroids could be responsible for both Q-type MBAs
and NEAs formation.
Acknowledgments
We also acknowledge the anonymous referees' useful suggestions for im-
proving the manuscript. This work is based on data collected at Subaru
Telescope, which is operated by the National Astronomical Observatory of
Japan. This work was supported in part by NSC Grant: NSC 101-2119-
M-008-007-MY3 and NSC 102-2119-M-008-001. The Pan-STARRS1 Surveys
(PS1) have been made possible through contributions by the Institute for
Astronomy, the University of Hawaii, the Pan-STARRS Project Office, the
Max-Planck Society and its participating institutes, the Max Planck Insti-
tute for Astronomy, Heidelberg and the Max Planck Institute for Extraterres-
trial Physics, Garching, The Johns Hopkins University, Durham University,
the University of Edinburgh, the Queen's University Belfast, the Harvard-
Smithsonian Center for Astrophysics, the Las Cumbres Observatory Global
Telescope Network Incorporated, the National Central University of Taiwan,
the Space Telescope Science Institute, and the National Aeronautics and
Space Administration under Grant No. NNX08AR22G issued through the
Planetary Science Division of the NASA Science Mission Directorate, the
National Science Foundation Grant No. AST-1238877, the University of
Maryland, Eotvos Lorand University (ELTE), and the Los Alamos National
Laboratory.
25
References
Bernstein, G., Khushalani, B. 2000. Orbit Fitting and Uncertainties for
Kuiper Belt Objects. The Astronomical Journal 120, 3323-3332.
Binzel, R. P., Xu, S., Bus, S. J., Skrutskie, M. F., Meyer, M. R., Knezek, P.,
Barker, E. S. 1993. Discovery of a Main-Belt Asteroid Resembling Ordinary
Chondrite Meteorites. Science 262, 1541-1543.
Binzel, R. P., Bus, S. J., Burbine, T. H., Sunshine, J. M. 1996. Spectral Prop-
erties of Near-Earth Asteroids: Evidence for Sources of Ordinary Chondrite
Meteorites. Science 273, 946-948.
Binzel, R. P., Harris, A. W., Bus, S. J., Burbine, T. H. 2001. Spectral Prop-
erties of Near-Earth Objects: Palomar and IRTF Results for 48 Objects
Including Spacecraft Targets (9969) Braille and (10302) 1989 ML. Icarus
151, 139-149.
Binzel, R. P., Rivkin, A. S., Stuart, J. S., Harris, A. W., Bus, S. J., Burbine,
T. H. 2004. Observed spectral properties of near-Earth objects: results for
population distribution, source regions, and space weathering processes.
Icarus 170, 259-294.
Binzel, R. P., Morbidelli, A., Merouane, S., DeMeo, F. E., Birlan, M., Ver-
nazza, P., Thomas, C. A., Rivkin, A. S., Bus, S. J., Tokunaga, A. T. 2010.
Earth encounters as the origin of fresh surfaces on near-Earth asteroids.
Nature 463, 331-334.
Bottke, W. F., Jr., Nolan, M. C., Greenberg, R. 1993. Collision lifetimes
and impact statistics of near-Earth asteroids. Lunar and Planetary Science
Conference 24, 159-160.
Bottke, W. F., Nolan, M. C., Greenberg, R., Kolvoord, R. A. 1994. Velocity
distributions among colliding asteroids. Icarus 107, 255-268.
Bottke, W. F., Morbidelli, A., Jedicke, R., Petit, J.-M., Levison, H. F.,
Michel, P., Metcalfe, T. S. 2002. Debiased Orbital and Absolute Magnitude
Distribution of the Near-Earth Objects. Icarus 156, 399-433.
Bottke, W. F., Durda, D. D., Nesvorn´y, D., Jedicke, R., Morbidelli, A.,
Vokrouhlick´y, D., Levison, H. 2005. The fossilized size distribution of the
main asteroid belt. Icarus 175, 111-140.
26
Bottke, W. F., Durda, D. D., Nesvorn´y, D., Jedicke, R., Morbidelli, A.,
Vokrouhlick´y, D., Levison, H. F. 2005. Linking the collisional history of
the main asteroid belt to its dynamical excitation and depletion. Icarus
179, 63-94.
Bowell, E., Chapman, C. R., Gradie, J. C., Morrison, D., Zellner, B. 1978.
Taxonomy of asteroids. Icarus 35, 313-335.
Bowell, E., Lumme, K. 1979. Colorimetry and magnitudes of asteroids. As-
teroids 132-169.
Brunetto, R., Vernazza, P., Marchi, S., Birlan, M., Fulchignoni, M., Orofino,
V., Strazzulla, G. 2006. Modeling asteroid surfaces from observations and
irradiation experiments: The case of 832 Karin. Icarus 184, 327-337.
Bus, S. J., Binzel, R. P. 2002. Phase II of the Small Main-Belt Asteroid
Spectroscopic Survey. The Observations. Icarus 158, 106-145.
Bus, S. J., Binzel, R. P. 2002. Phase II of the Small Main-Belt Asteroid
Spectroscopic Survey. A Feature-Based Taxonomy. Icarus 158, 146-177.
Carvano, J. M., Hasselmann, P. H., Lazzaro, D., Moth´e-Diniz, T. 2010.
SDSS-based taxonomic classification and orbital distribution of main belt
asteroids. Astronomy and Astrophysics 510, AA43.
Chang, C.-K., Waszczak, A., Lin, H.-W., Ip, W.-H., Prince, T. A., Kulka-
rni, S. R., Laher, R., Surace, J. 2014. A New Large Super-fast Rotator:
(335433) 2005 UW163. The Astrophysical Journal 791, L35.
Chapman, C. R. 1996. S-Type Asteroids, Ordinary Chondrites, and Space
Weathering: The Evidence from Galileo's Fly-bys of Gaspra and Ida. Me-
teoritics and Planetary Science 31, 699-725.
Chapman, C. R. 2004. Space Weathering of Asteroid Surfaces. Annual Re-
view of Earth and Planetary Sciences 32, 539-567.
Chapman, C. R. 2010. Asteroids: Stripped on passing by Earth. Nature 463,
305-306.
Clark, B. E., and 11 colleagues 2001. Space weathering on Eros: Constraints
from albedo and spectral measurements of Psyche crater. Meteoritics and
Planetary Science 36, 1617-1637.
27
Clark, B. E., Helfenstein, P., Bell, J. F., Peterson, C., Veverka, J., Izenberg,
N. I., Domingue, D., Wellnitz, D., McFadden, L. 2002. NEAR Infrared
Spectrometer Photometry of Asteroid 433 Eros. Icarus 155, 189-204.
Dandy, C. L., Fitzsimmons, A., Collander-Brown, S. J. 2003. Optical colors
of 56 near-Earth objects: trends with size and orbit. Icarus 163, 363-373.
Davis, D. R., Durda, D. D., Marzari, F., Campo Bagatin, A., Gil-Hutton,
R. 2002. Collisional Evolution of Small-Body Populations. Asteroids III
545-558.
Delbo, M., Libourel, G., Wilkerson, J., Murdoch, N., Michel, P., Ramesh,
K. T., Ganino, C., Verati, C., Marchi, S. 2014. Thermal fatigue as the
origin of regolith on small asteroids. Nature 508, 233-236.
DeMeo, F. E., Binzel, R. P., Slivan, S. M., Bus, S. J. 2009. An extension of
the Bus asteroid taxonomy into the near-infrared. Icarus 202, 160-180.
DeMeo, F. E., Carry, B. 2013. The taxonomic distribution of asteroids from
multi-filter all-sky photometric surveys. Icarus 226, 723-741.
DeMeo, F. E., Carry, B. 2014. Solar System evolution from compositional
mapping of the asteroid belt. Nature 505, 629-634.
DeMeo, F. E., Binzel, R. P., Lockhart, M. 2014. Mars encounters cause fresh
surfaces on some near-Earth asteroids. Icarus 227, 112-122.
Hapke, B. 2001. Space weathering from Mercury to the asteroid belt. Journal
of Geophysical Research 106, 10039-10074.
Holmberg, J., Flynn, C., Portinari, L. 2006. The colours of the Sun. Monthly
Notices of the Royal Astronomical Society 367, 449-453.
Holsapple, K. A. 2007. Spin limits of Solar System bodies: From the small
fast-rotators to 2003 EL61. Icarus 187, 500-509.
Ishiguro, M., and 18 colleagues 2007. Global mapping of the degree of space
weathering on asteroid 25143 Itokawa by Hayabusa/AMICA observations.
Meteoritics and Planetary Science 42, 1791-1800.
28
Ivezi´c, Z., and 32 colleagues 2001. Solar System Objects Observed in the
Sloan Digital Sky Survey Commissioning Data. The Astronomical Journal
122, 2749-2784.
Jacobson, S. A., Scheeres, D. J. 2011. Dynamics of rotationally fissioned
asteroids: Source of observed small asteroid systems. Icarus 214, 161-178.
Jedicke, R., Nesvorn´y, D., Whiteley, R., Ivezi´c, Z., Juri´c, M. 2004. An age-
colour relationship for main-belt S-complex asteroids. Nature 429, 275-277.
Lantz, C., Clark, B. E., Barucci, M. A., Lauretta, D. S. 2013. Evidence for
the effects of space weathering spectral signatures on low albedo asteroids.
Astronomy and Astrophysics 554, A138.
Lazzaro, D., Angeli, C. A., Carvano, J. M., Moth´e-Diniz, T., Duffard, R.,
Florczak, M. 2004. S 3OS 2: the visible spectroscopic survey of 820 aster-
oids. Icarus 172, 179-220.
Magnier, E. A., and 13 colleagues 2013. The Pan-STARRS 1 Photometric
Reference Ladder, Release 12.01. The Astrophysical Journal Supplement
Series 205, 20.
Marchi, S., Magrin, S., Nesvorn´y, D., Paolicchi, P., Lazzarin, M. 2006. A
spectral slope versus perihelion distance correlation for planet-crossing as-
teroids. Monthly Notices of the Royal Astronomical Society 368, L39-L42.
Migliorini, F., Michel, P., Morbidelli, A., Nesvorny, D., Zappala, V. 1998.
Origin of Multikilometer Earth- and Mars-Crossing Asteroids: A Quanti-
tative Simulation. Science 281, 2022.
Miyazaki, S., and 14 colleagues 2002. Subaru Prime Focus Camera -- Suprime-
Cam. Publications of the Astronomical Society of Japan 54, 833-853.
Morbidelli, A., Vokrouhlick´y, D. 2003. The Yarkovsky-driven origin of near-
Earth asteroids. Icarus 163, 120-134.
Morbidelli, A., Bottke, W. F., Nesvorn´y, D., Levison, H. F. 2009. Asteroids
were born big. Icarus 204, 558-573.
Moth´e-Diniz, T., Nesvorn´y, D. 2008. Visible spectroscopy of extremely young
asteroid families. Astronomy and Astrophysics 486, L9-L12.
29
Nakamura, T., Yoshida, F. 2002. Statistical Method for Deriving Spatial and
Size Distributions of Sub-km Main-Belt Asteroids from Their Sky Motions.
Publications of the Astronomical Society of Japan 54, 1079-1089.
Nesvorn´y, D., Jedicke, R., Whiteley, R. J., Ivezi´c, Z. 2005. Evidence for
asteroid space weathering from the Sloan Digital Sky Survey. Icarus 173,
132-152.
Nesvorn´y, D., Bottke, W. F., Vokrouhlick´y, D., Chapman, C. R., Rafkin,
S. 2010. Do planetary encounters reset surfaces of near Earth asteroids?.
Icarus 209, 510-519.
Padmanabhan, N., and 23 colleagues 2008. An Improved Photometric Cali-
bration of the Sloan Digital Sky Survey Imaging Data. The Astrophysical
Journal 674, 1217-1233.
Petit, J.-M., Holman, M. J., Gladman, B. J., Kavelaars, J. J., Scholl, H.,
Loredo, T. J. 2006. The Kuiper Belt luminosity function from mR= 22 to
25. Monthly Notices of the Royal Astronomical Society 365, 429-438.
Polishook, D., Moskovitz, N., Binzel, R. P., DeMeo, F. E., Vokrouhlick´y,
D., Zizka, J., Oszkiewicz, D. 2014. Observations of fresh and weathered
surfaces on asteroid pairs and their implications on the rotational-fission
mechanism. Icarus 233, 9-26.
Rabinowitz, D. L. 1997. Are Main-Belt Asteroids a Sufficient Source for the
Earth-Approaching Asteroids?. Icarus 127, 33-54.
Rivkin, A. S., Thomas, C. A., Trilling, D. E., Enga, M.-t., Grier, J. A. 2011.
Ordinary chondrite-like colors in small Koronis family members. Icarus
211, 1294-1297.
Rozitis, B., Maclennan, E., Emery, J. P. 2014. Cohesive forces prevent the
rotational breakup of rubble-pile asteroid (29075) 1950 DA. Nature 512,
174-176.
Rubincam, D. P. 2000. Radiative Spin-up and Spin-down of Small Asteroids.
Icarus 148, 2-11.
Sasaki, S., Nakamura, K., Hamabe, Y., Kurahashi, E., Hiroi, T. 2001. Pro-
duction of iron nanoparticles by laser irradiation in a simulation of lunar-
like space weathering. Nature 410, 555-557.
30
Schlafly, E. F., and 16 colleagues 2012. Photometric Calibration of the First
1.5 Years of the Pan-STARRS1 Survey. The Astrophysical Journal 756,
158.
Scheeres, D. J. 2007. The dynamical evolution of uniformly rotating asteroids
subject to YORP. Icarus 188, 430-450.
Scheeres, D. J. 2015. Landslides and Mass shedding on spinning spheroidal
asteroids. Icarus 247, 1-17.
Stuart, J. S., Binzel, R. P. 2004. Bias-corrected population, size distribution,
and impact hazard for the near-Earth objects. Icarus 170, 295-311.
Tholen, D. J. 1984. Asteroid taxonomy from cluster analysis of Photometry.
Ph.D. Thesis .
Thomas, C. A., Rivkin, A. S., Trilling, D. E., Enga, M.-t., Grier, J. A. 2011.
Space weathering of small Koronis family members. Icarus 212, 158-166.
Thomas, C. A., Trilling, D. E., Rivkin, A. S. 2012. Space weathering of small
Koronis family asteroids in the SDSS Moving Object Catalog. Icarus 219,
505-507.
Tonry, J. L., and 13 colleagues 2012. The Pan-STARRS1 Photometric Sys-
tem. The Astrophysical Journal 750, 99.
Veres, P., Jedicke, R., Denneau, L., Wainscoat, R., Holman, M. J., Lin, H.-W.
2012. Improved Asteroid Astrometry and Photometry with Trail Fitting.
Publications of the Astronomical Society of the Pacific 124, 1197-1207.
Vernazza, P., Binzel, R. P., Rossi, A., Fulchignoni, M., Birlan, M. 2009.
Solar wind as the origin of rapid reddening of asteroid surfaces. Nature
458, 993-995.
Walsh, K. J., Richardson, D. C., Michel, P. 2012. Spin-up of rubble-pile
asteroids: Disruption, satellite formation, and equilibrium shapes. Icarus
220, 514-529.
Willman, M., Jedicke, R., Nesvorn´y, D., Moskovitz, N., Ivezi´c, Z., Fevig,
R. 2008. Redetermination of the space weathering rate using spectra of
Iannini asteroid family members. Icarus 195, 663-673.
31
Willman, M., Jedicke, R., Moskovitz, N., Nesvorn´y, D., Vokrouhlick´y, D.,
Moth´e-Diniz, T. 2010. Using the youngest asteroid clusters to constrain
the space weathering and gardening rate on S-complex asteroids. Icarus
208, 758-772.
Willman, M., Jedicke, R. 2011. Asteroid age distributions determined by
space weathering and collisional evolution models. Icarus 211, 504-510.
Yoshida, F., Nakamura, T. 2007. Subaru Main Belt Asteroid Survey (SM-
BAS) - Size and color distributions of small main-belt asteroids. Planetary
and Space Science 55, 1113-1125.
Zellner, B., Tholen, D. J., Tedesco, E. F. 1985. The eight-color asteroid
survey - Results for 589 minor planets. Icarus 61, 355-416.
32
|
1603.09391 | 2 | 1603 | 2016-11-02T19:16:05 | New Planetary Systems from the Calan-Hertfordshire Extrasolar Planet Search | [
"astro-ph.EP",
"astro-ph.SR"
] | We report the discovery of eight new giant planets, and updated orbits for four known planets, orbiting dwarf and subgiant stars using the CORALIE, HARPS, and MIKE instruments as part of the Calan-Hertfordshire Extrasolar Planet Search. The planets have masses in the range 1.1-5.4MJs, orbital periods from 40-2900 days, and eccentricities from 0.0-0.6. They include a double-planet system orbiting the most massive star in our sample (HD147873), two eccentric giant planets (HD128356b and HD154672b), and a rare 14 Herculis analogue (HD224538b). We highlight some population correlations from the sample of radial velocity detected planets orbiting nearby stars, including the mass function exponential distribution, confirmation of the growing body of evidence that low-mass planets tend to be found orbiting more metal-poor stars than giant planets, and a possible period-metallicity correlation for planets with masses >0.1MJ, based on a metallicity difference of 0.16 dex between the population of planets with orbital periods less than 100 days and those with orbital periods greater than 100 days. | astro-ph.EP | astro-ph | MNRAS 000, 1–20 (2002)
Preprint 3 November 2016
Compiled using MNRAS LATEX style file v3.0
New Planetary Systems from the Calan-Hertfordshire
Extrasolar Planet Search
J.S. Jenkins1⋆, H.R.A. Jones2, M. Tuomi2, M. D´ıaz1, J.P. Cordero1, A. Aguayo1,
B. Pantoja1, P. Arriagada3, R. Mahu4, R. Brahm5, P. Rojo1, M.G. Soto1, O. Ivanyuk6,
N. Becerra Yoma4, A.C. Day-Jones1, M.T. Ruiz1, Y.V. Pavlenko2,6, J.R. Barnes7,
F. Murgas8, D.J. Pinfield2, M.I. Jones9, M. L´opez-Morales10, S. Shectman11,
R.P. Butler3, D. Minniti12
1Departamento de Astronomia, Universidad de Chile, Camino el Observatorio 1515, Las Condes, Santiago, Chile, Casilla 36-D
2Center for Astrophysics, University of Hertfordshire, College Lane Campus, Hatfield, Hertfordshire, UK, AL10 9AB
3Carnegie Institution of Washington, Dept. of Terrestrial Magnetism, 5241 Broad Branch Rd. NW, 20015, Washington D.C., USA
4Dept. of Electrical Engineering, Universidad de Chile, Av. Tupper 2007, PO Box 412-3, Santiago, Chile
5Instituto de Astrof´ısica, Facultad de F´ısica, Pontificia Universidad Cat´olica de Chile, Av. Vicuna Mackenna 4860, Santiago, Chile
6Main Astronomical Observatory, Academy of Sciences of Ukraine, Golosiiv Woods, Kyiv-127, 03680 Ukraine
7Department of Physical Sciences, The Open University, Walton Hall, Milton Keynes, MK7 6AA, UK
8Instituto de Astrof´ısica de Canarias, Via Lactea, E38205, La Laguna, Tenerife, Spain
9Center of Astro-Engineering UC, Pontificia Universidad Cat´olica de Chile, Av. Vicuna Mackenna 4860, Santiago, Chile
10Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138
11Carnegie Institution of Washington, The Observatories, 813 Santa Barbara Street, Pasadena, CA 91101-1292, USA
12Departamento de Ciencias Fisicas, Universidad Andres Bello, Republica 220, Santiago, Chile
Submitted January 2011
ABSTRACT
We report the discovery of eight new giant planets, and updated orbits for four
known planets, orbiting dwarf and subgiant stars using the CORALIE, HARPS, and
MIKE instruments as part of the Calan-Hertfordshire Extrasolar Planet Search. The
planets have masses in the range 1.1-5.4MJ's, orbital periods from 40-2900 days, and
eccentricities from 0.0-0.6. They include a double-planet system orbiting the most
massive star in our sample (HD147873), two eccentric giant planets (HD128356b and
HD154672b), and a rare 14 Herculis analogue (HD224538b). We highlight some popu-
lation correlations from the sample of radial velocity detected planets orbiting nearby
stars, including the mass function exponential distribution, confirmation of the grow-
ing body of evidence that low-mass planets tend to be found orbiting more metal-poor
stars than giant planets, and a possible period-metallicity correlation for planets with
masses >0.1 MJ, based on a metallicity difference of 0.16 dex between the popula-
tion of planets with orbital periods less than 100 days and those with orbital periods
greater than 100 days.
Key words:
stars: planetary systems; planets and satellites: formation; stars: activity; stars: low-
mass; stars: solar-type
6
1
0
2
v
o
N
2
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
9
3
9
0
.
3
0
6
1
:
v
i
X
r
a
1
INTRODUCTION
After the discovery and confirmation of the hot Jupiter
Dimidium (aka. Helvetios b or 51 Pegasi b) in 1995, our view
of giant planets was changed forever. Giant planets orbiting
⋆ E-mail: [email protected]
c(cid:13) 2002 The Authors
dwarf stars like the Sun have been found to inhabit many re-
gions of the parameter space. The first of these were found
orbiting their stars with periods much shorter than those
of Jupiter in our solar system (e.g. Mayor & Queloz 1995;
Marcy & Butler 1996). More recently as the number of giant
planets has grown, a new population of eccentric gas giants
has been shown to exist (e.g. Jones et al. 2006; Tamuz et al.
2008; Arriagada et al. 2010), making up a large fraction of
2
J.S. Jenkins et al.
the known systems, at least around main sequence stars
since the same high fraction does not appear to be present
around giant stars (Jones et al. 2014). These planets seem
to span the full range of masses from the sub-Jupiter range
all the way up to the mass boundary between giant planets
and brown dwarfs.
It has been thoroughly demonstrated that there is
a clear bias to metal-rich stars hosting giant planets
across the currently sampled range of orbital separa-
tions (Fischer & Valenti 2005; Sousa et al. 2011), a bias
that seems to hold for stars beyond the main sequence
(e.g. Reffert et al. 2015; Soto et al. 2015; Jones et al. 2016).
In comparison, recent work appears to show that stars
that host lower-mass planets have a metallicity distribu-
tion that is indistinguishable from those that host no
known planets at all (Udry et al. 2007; Jenkins et al. 2013b;
Sousa et al. 2011; Buchhave & Latham 2015). Furthermore,
Jenkins et al. suggests there exists a boundary that delimits
a planet desert for the lowest mass planets in the metal-rich
regime. Such results show there is a fundamental relation-
ship between the proto-planetary disc metallicity and the
mass of planets that form in these discs, indicating further
study of the planetary mass function and its relationship
with metallicity is warranted.
Since metallicity plays a key role in the formation
of the observed planetary systems, explained well by na-
ture's merging of core accretion and planet migration (e.g.
Ida & Lin 2004; Mordasini et al. 2012), and now we are
reaching a population size where less obvious correlations
can reveal themselves, studying the current population of
known planets can provide a window into the fundamentals
of planet formation and evolution for the stars nearest to
the Sun.
In 2007 we started a radial velocity planet search pro-
gram on the HARPS instrument at La Silla Chile, with
the goal of finding more gas giant planets orbiting super
metal-rich stars to increase the statistics, whilst also fol-
lowing up any discoveries to search for transit events. The
first results from this work and the target sample were
discussed in Jenkins et al. (2009) and since then our pro-
gram has expanded to make use of the CORALIE spectro-
graph (Jenkins et al. 2011b,a, 2013a). This current paper
announces the first planets from this survey detected using
CORALIE and goes on to study the planet mass-metallicity
plane to search for correlations between these two parame-
ters.
In this work we describe the latest efforts from our
Calan-Hertfordshire Extrasolar Planet Search, which builds
on previous work by this group. We include the new giant
planets we have detected in this program with the large sam-
ple of gas giants detected by radial velocity measurements
that already exist in the literature, in order to continue the
search for emerging correlations that allow us a more strin-
gent insight into the nature of these objects. In § 2 we de-
scribe the measurements used in this work. In § 3 we discuss
the sample selection, introduce the new giant planet detec-
tions from this survey, and discuss some characteristics of
their host stars. In § 4 we perform tests of the mass function
and its relationship to stellar metallicity, along with search-
ing for correlations between these parameters and planetary
orbital period and we briefly discuss the impact of these re-
sults. Finally, in § 5 we summerise the main points of the
present work.
2 OBSERVATIONS AND REDUCTION
The radial velocity datasets for these stars were observed
using the precision radial velocity spectrographs CORALIE,
HARPS, and MIKE. Both the CORALIE and HARPS spec-
trographs are physically located at the ESO La Silla Obser-
vatory in Chile, where CORALIE is mounted on the Swiss
Euler telescope and HARPS is fed by light from the ESO
3.6m telescope. The MIKE spectrograph is located at the
Las Campanas Observatory and is mounted on the Magel-
lan Clay 6.5m telescope.
In this work, 570 radial velocities are reported, with a
fairly even split between CORALIE and HARPS observa-
tions, in comparison to the smaller fraction of MIKE data.
The baseline of observations for the CORALIE data run
from 2009 November 25th until 2015 October 23rd (BJD
2455160.53623 - 2457318.85147), whereas the HARPS data
runs from 2007 May 28th until 2013 September 28th (BJD
2454248.60231 - 2456563.90982) showing that the HARPS
data covers a longer baseline but the CORALIE data has
better sampling coverage in general for these targets. The
MIKE velocities run from 2003 August 13th until 2009 July
7th (BJD 2452864.57934 - 2455019.6938), covering a base-
line of six years that overlaps with the HARPS baseline but
not the CORALIE data.
2.1 CORALIE
The analysis of CORALIE data involves the normal echelle
reduction steps, such as debiasing the images using CCD
bias frames, order location and tracing using polynomial
fitting methods and aperture order filtering, pixel-to-pixel
sensitivity correction (flatfielding) by building a normalised
master flatfield image and dividing out the master flat-
field from the other images, scattered-light removal by mea-
suring the contribution to the total
light profile in the
inter-order regions, order extraction using a profile fitting
method (Marsh 1989), and finally building a precise 2D
wavelength solution that is good to ∼2.5 m s−1overall preci-
sion (Jord´an et al. 2014; Brahm & Jordan 2016). The final
steps in the analysis include cross-correlating the individual
spectra with a binary mask (either G2, K0, K5, or M2 de-
pending on the spectral type of the star, see Baranne et al.
1996 and Pepe et al. 2002) and fitting the cross-correlation
function (CCF) with a gaussian to measure the radial ve-
locity, along with the width of the CCF to generate realis-
tic uncertainties. The instrumental drift is then measured
and removed from the overall velocity by performing a sec-
ond cross-correlation between the simultaneously measured
Thorium-Argon (ThAr) lamp (simultaneously referring to
observations where the second optical fibre illuminates the
spectrograph with ThAr light) and a previously measured
double ThAr observation where both fibres have been fed
by light from the ThAr calibration lamp. These double
ThAr measurements are generally taken every 1.5-2 hours
throughout the night to continually reset the wavelength so-
lution zero-point. These steps were discussed in Jord´an et al.
MNRAS 000, 1–20 (2002)
(2014), however the overall performance in term of stability
is shown in Appendix A.
An important additional step in the calculation of the
radial velocities from CORALIE is the characterisation of
the offset between the data collected prior to the November
2014 upgrade of the instrument. As part of this upgrade, the
CORALIE circular fibres were replaced with octagonal fibres
to increase illumination stability, the double-scrambler was
reintroduced into the system, and a focal mirror that focuses
light on the guiding camera was replaced. Such instrumen-
tal upgrades are expected to introduce systematic offsets in
the radial velocity measurements compared to pre-upgrade
observations.
In an attempt to account for the offset between the
CORALIE reference velocities observed before and after the
upgrade, we first determined this offset in our two radial
velocity reference targets, HD72673 and HD157347. We de-
note the mean estimate for the offset x0 based on these
reference targets as µ0 and its standard error as σ0. We
then used these numbers to construct a prior probability for
the offset x1 in the first data set of our sample such that
π(x1) = N (µ0, σ2
0). This prior was used to calculate a pos-
terior for the offset and we denote the mean and standard
deviation of this obtained posterior with µ1 and σ1, respec-
tively. After that, we adopted the refined estimate of the
offset of µi ± σi to construct the prior for the offset xi+1
in the next data set, such that π(xi+1) = N (µi, σ2
i ). This
process, called 'Bayesian updating' because the prior is up-
dated into a posterior density that is in turn used as as the
next prior, was repeated until all the data sets were anal-
ysed. This process helps to account for all the information
regarding the offsets from all the data sets without having
to analyse them simultaneously. As a result, we summarise
the information regarding the offset as a probability distri-
bution that is almost Gaussian in the sense that the third
and fourth moments are very close to zero. This density has
a mean of 19.2 ms−1 an a standard deviation of 4.8 ms−1
, which indicates that an offset in the reference velocity of
roughly 20 ms−1 is significantly present in the CORALIE
data sets after the upgrade for the stars included in this
work.
2.2 HARPS
For HARPS, the steps are similar to those mentioned above,
but the data is automatically processed by the HARPS-DRS
version 3.5 which is based in general on the procedure ex-
plained in Baranne et al. (1996). The nightly drift of the
ThAr lines are found to be below 0.5 m s−1and including
the other sources of uncertainty such as centering and guid-
ing (<30 cm s−1), a stability of less than 1 m s−1is found
for this spectrograph over the long term (see Lo Curto et al.
2010).
2.3 MIKE
For the radial velocities measured using the MIKE spectro-
graph, the reduction steps are similar but the analysis pro-
cedure is different. MIKE uses a cell filled with molecular
iodine (I2) that is placed directly in the beam of light from
the target star before entering the spectrograph, and this is
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
3
used to record the instrument point spread function (PSF)
and provide a highly accurate wavelength fiducial. The full
analysis procedure is explained in Butler et al. (1996), but
in short, the velocities are measured by comparing each
star+I2 spectrum to that of a template measurement of the
same star. This template is observed without the I2 cell in
place, such that one can deconvolve the instrument's PSF
from the template observation, usually accomplished by ob-
serving a rapidly rotating B-star with the I2 cell in place
before and after the template observation and extrapolating
the PSF from these observations to that of the template. The
deconvolved template can then be used to forward model
each observation by convolving it with a very high resolu-
tion and high S/N I2 spectrum and a modeled PSF. The final
stability of MIKE is found to be between that of CORALIE
and HARPS, around the 5 m s−1level of velocity precision.
3 THE STARS AND THEIR DOPPLER
SIGNALS
The selection of the Calan-Hertfordshire Extrasolar Planet
Search (CHEPS) target sample is discussed in Jenkins et al.
(2009) but we outline the main selection criteria here. The
stars are generally selected to be late-F to early-K stars,
with a B − V range between 0.5-0.9, and with a small num-
ber of redder stars included to allow the study of activ-
ity correlations and timescales between activity indicators
and the measured Doppler velocities. The positions of the
CHEPS sample on a colour-magnitude diagram are shown
in Fig. 1 (squares) compared to a general selection of nearby
stars from the Hipparcos catalogue (circles), along with the
targets discussed in this work (stars). The magnitude lim-
its we set are between 7.5-9.5 in the V optical band and
after preliminary screening with the FEROS spectrograph
(Jenkins et al. 2008), we predominantly selected chromo-
spherically quiet stars, and those that are metal-rich (pri-
mary sample having logR′
HK indices 6 -4.9 dex and [Fe/H] >
+0.1 dex, with some stars outside of these selection limits to
use as comparisons). The positions of the stars reported here
on an HR-diagram are shown in the inset plot in Fig. 1. The
Y2 isomass tracks are also shown (Demarque et al. 2004)
for masses of 1.0-1.3 M⊙ , increasing in mass towards higher
luminosities, and with a fixed metallicity of +0.2 dex. The
characteristics of the targets we discuss in this work are
shown in Table 1.
Since the planets reported here are gas giants, the ra-
dial velocity signals of these stars are fairly large, and hence
one might expect that we can detect frequencies for all of
them using standard periodogram analyses to hunt for power
peaks in the Fourier power spectrum, or minimum mean
square error (MMSE) spectrum (e.g. Dawson & Fabrycky
2010; Jenkins et al. 2014). However, the search for the best
solutions can be complicated, particularly for combined data
sets from independent spectrographs because the sets have
different baselines, data samplings, and uncertainties that
all have to be accounted for in a search for periodic signals.
The signals we discuss in this work are a mixture of well,
and moderately well sampled data, such that some signals
are more problematic to detect than others. Moreover, the
inclusion of a linear trend that could be present in the data
of a given target due to a long period companion, will cause
4
J.S. Jenkins et al.
Table 1. Stellar parameters for the hosts.
Parameter
HD9174
HD48265
HD68402
HD72892
HD128356
HD143361
′
HK
Spectral TypeHipp
B − VHipp
V
π (mas)
distance (pc)
logR
Hipparcos N obs
Hipparcos σ
∆MV
T eff (K)
L⋆/L⊙
M⋆ /M⊙
R⋆/R⊙
[Fe/H]
log g
U,V,W (km/s)
vsini (km/s)
Age (Gyrs)
G8IV
0.761±0.002
8.40
12.67±0.62
78.93±3.86
-5.23
92
0.014
1.478
5577±100
2.41±0.18
1.03±0.05
1.67±0.07
0.39±0.10
4.03±0.05
22.2,-56.5,-29
2.1±0.2
9±3
G5IV/V
0.747±0.014
8.05
11.71±0.58
85.40±4.23
-5.24
98
0.009
1.914
5650±100
3.84±0.19
1.28±0.05
2.05±0.05
0.40±0.10
3.92±0.03
-14.2,-24.0,4.5
3.1±0.3
5±3
G5IV/V
0.660±0.021
9.11
12.82±0.61
78.00±3.71
-4.95
107
0.017
0.108
5950±100
1.17±0.06
1.12±0.05
1.02±0.05
0.29±0.10
4.47±0.05
-37.5,-16.3,-17.3
2.9±0.2
2±3
G5V
0.686±0.015
8.83
13.74±0.83
72.78±4.40
-5.03
92
0.016
0.405
5688±100
1.40±0.09
1.02±0.05
1.22±0.06
0.25±0.10
4.27±0.05
72.2,-2.0,-15.8
2.7±0.2
8±3
K3V
1.017±0.015∗
8.29
38.41±0.77
26.03±0.52
-5.07
72
0.013
0.553
4875±100
0.36±0.01
0.65±0.05
0.85±0.02
0.17±0.10
4.52±0.06
30.8,-28.7,8.8
1.3±0.1
10±5
G6V
0.773±0.004
9.20
15.23±1.18
65.66±5.09
-5.12
96
0.016
0.349
5505±100
0.81±0.06
0.95±0.05
0.99±0.08
0.22±0.10
4.42±0.08
-24.4,-49.5,3.7
1.5±0.1
5±5
Parameter
HD147873
HD152079
HD154672
HD165155
HD224538
′
HK
Spectral TypeHip
B − VHip
V
π (mas)
distance (pc)
logR
Hipparcos N obs
Hipparcos σ
∆MV
T eff (K)
L⋆/L⊙
M⋆ /M⊙
R⋆/R⊙
[Fe/H]
log g
U,V,W (km/s)
vsini (km/s)
Age (Gyrs)
G1V
0.575±0.012
7.96
9.53±0.99
104.93±10.90
-5.27
88
0.011
1.337
5972±100
5.99±0.62
1.38±0.05
2.29±0.10
-0.03±0.10
3.86±0.05
18.7,-18.2,3.2
5.9±0.5
4±3
G6V
0.711±0.025
9.18
12.00±1.52
83.33±10.56
-4.99
84
0.014
0.508
5726±100
1.28±0.16
1.10±0.05
1.15±0.13
0.16±0.10
4.36±0.10
-39.6,-46.2,10.2
1.8±0.1
3±3
G3IV
0.713±0.013
8.21
15.44±0.84
64.77±3.52
-5.12
120
0.014
0.943
5655±100
1.91±0.10
1.08±0.05
1.44±0.05
0.11±0.10
4.15±0.05
-20.0,-18.7,-29.1
2.2±0.2
8±3
G8V
1.018±0.095
9.36
15.39±1.72
64.98±7.26
-5.18
67
0.019
1.373
5426±100
0.70±0.08
1.02±0.05
0.95±0.11
0.09±0.10
4.49±0.11
13.3,9.8,-20.5
1.5±0.1
11±4
F9IV/V
0.581±0.006
8.06
12.86±0.73
77.76±4.41
-4.99
163
0.017
0.627
6097±100
2.95±0.17
1.34±0.05
1.54±0.06
0.27±0.10
4.19±0.04
-29.1,-15.0,+7.2
3.9±0.3
2±3
* - colour calculated from magnitudes drawn from the Tycho-2 catalogue (Høg et al. 2000).
We assign a standard ±100K uncertainty to the Teff measurements and ±0.10 dex to the metallicities.
π values come from van Leeuwen (2007) and all other Hipparcos parameters are taken from Perryman et al. (1997).
Evolutionary bulk properties and spectrally measured indices were either computed in this work or taken from Jenkins et al. (2008), Jenkins et al. (2011b), and
Murgas et al. (2013).
[Fe/H] abundances, spectroscopic log g values, and rotational velocities were calculated using the procedures in Pavlenko et al. (2012) and taken from Ivanyuk et al.
(2016).
Figure 1. Positions of these stars on a colour magnitude diagram. The grey filled circles are nearby stars drawn from the Hipparcos
catalogue, the filled squares are the full sample of CHEPS targets, the filled stars are the positions of the stars discussed in this work,
and the solid curve marks the position of the main sequence. In the inset we show an HR-diagram with the positions of the stars
discussed here, where the Y2 evolutionary models for masses of 1.0 (solid), 1.1 (dashed), 1.2 (dot-dashed), and 1.3 (dot-dot-dashed) and
metallicities of +0.2 dex are shown increasing towards higher luminosities and temperatures. The arrow shows the approximate direction
of increasing mass.
MNRAS 000, 1–20 (2002)
considerable correlations to the probability densities of the
model parameters.
In Appendix B1 we show the MMSE power spectra for
our sample of stars and it is clear that a number of deep
troughs are present in the data, yet some sets show no sig-
nificant power troughs over the frequency space searched.
We thus applied the delayed-rejection adaptive-Metropolis
algorithm (DRAM; Haario et al. 2006) together with tem-
pered Markov chains when searching for solutions to our
Keplerian models (see Tuomi 2014; Tuomi et al. 2014) and
the simpler adaptive-Metropolis algorithm (Haario et al.
2001) when obtaining estimates for the model parameters.
This method has previously been applied in Tuomi et al.
(2013), Jenkins et al. (2013b), Jenkins et al. (2013c), and
Jenkins & Tuomi (2014), for example. The vertical dashed
lines in the MMSE periodograms mark the positions of these
detected signals.
The DRAM algorithm works by using a sequence of pro-
posal densities that are each narrower than the last. Here
"narrower" is to be understood as a multivariate Gaussian
proposal density, on which the adaptive-Metropolis algo-
rithm is based, that has a smaller variance for at least one of
the parameters in the parameter vector. In short, if a value
proposed by drawing it from an initial proposal density is re-
jected, we continue by modifying the proposal density with
respect to the period parameter(s) by multiplying it with a
factor of 0.1 such that another value is proposed from a nar-
rower area surrounding the current state of the chain. This
enables us to visit the areas of high probability in the pe-
riod space repeatedly and reliably and to see which periods
correspond to the highest values of the posterior probability
density.
We applied a statistical model that contains reference
velocities of each instrument, a linear trend, Keplerian sig-
nals, and excess white noise. By simplifying the statistical
model in Tuomi et al. (2014) by removing the intrinsic cor-
relations that we do not expect to play a significant role due
to the low number of measurements and the (relatively) high
amplitude variations in the data, the model can be written
as
mi,l = γl + γti + fk(ti) + ǫi,l
(1)
l + σ2
where mi,l
is the ith measurement made by the lth
telescope/instrument, γl is the systemic velocity offset for
each instrument,
γ represents the linear trend, fk is the
superposition of k Keplerian signals at time ti, and ǫi,l is
a Gaussian random variable with zero mean and a vari-
ance of σ2
i , where σl is a free parameter in our anal-
yses that represents the stellar jitter noise and σi is the
measurement uncertainty. We use prior probability densi-
ties as described in Tuomi (2012) but apply an informa-
tive prior density for the orbital eccentricities such that
π(e) ∝ N (0, 0.32) that penalises, but does not exclude,
eccentricities close to unity (Anglada-Escud´e et al. 2013;
Tuomi & Anglada-Escud´e 2013). Therefore, given the pre-
vious tests performed using this model, in particular using
high-precision HARPS data, we expect this model to repre-
sent a sufficiently accurate description of the velocities anal-
ysed in the current work. The good agreement between the
parameters published for the stars we discuss that overlap
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
5
Figure 2. Phase folded CORALIE (blue) and HARPS (red) ve-
locities for HD9174. The solid curve is the best fit Keplerian
model.
with previous work, is testament to the model's applicabil-
ity.
Posterior densities for the periods, semi-amplitudes,
and eccentricities for all signals are shown in Appen-
dices C1, C2, and C3, and below we discuss each of the sys-
tems independently. Note that all the signal fits are shown
phase folded to highlight the phase coverage we have ob-
served. We did not show the posteriors for the residuals to
any of the systems (more than a single planet model fit) un-
less we detected a unique solution that relates to a second
Doppler signal.
3.1 HD9174
HD9174 is classed as a G8 subgiant star in the Hipparcos
catalogue (Perryman et al. 1997) since it has a distance from
the main sequence (∆MV ) of 1.5 and a B − V colour of
0.76. The star is also extremely metal-rich, with an [Fe/H]
of nearly +0.4 dex, indicating a high probability of hosting a
giant planet. A signal with a semi-amplitude of 21 m s−1has
been detected with a period of nearly 1200 days. The star is
very chromospherically quiet and exhibits low rotational ve-
locity, indicating it is an ideal subgiant to search for orbiting
exoplanets using the radial velocity method, and hence we
do not believe the signal is induced by activity features, and
therefore conclude that HD9174 hosts an orbiting planet.
In order to understand the significance of the detected
signals we present here, we calculated log-Bayesian evidence
ratios for each of the signals presented here based on the
MCMC samples drawn from a mixture of prior and posterior
densities, as discussed in Newton & Raftery (1994) (see Eqs.
15 & 16). This is a version of importance sampling that uses
a mixed distribution to obtain a sampling distribution that
has "heavier" tails than the posterior (see Nelson et al. 2016
for further discussion of this method). We note that all of
these signals are so strong that they would pass essentially
any meaningful significance test, from likelihood-ratio tests
to tests applying any other information criteria.
The log-Bayesian evidence ratio for the signal in the
data of HD9174 was found to be 42.1 for the 1-planet
model. The signal is also readily apparent in the MMSE pe-
riodogram. No secondary signal was detected. Considering
6
J.S. Jenkins et al.
Figure 3. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for the long period signal detected in
the HD48265 velocities. The solid curve represents the best fit
Keplerian model to the data.
Figure 4. Phase folded CORALIE (blue) and HARPS (red) ve-
locities for HD68402. The solid curve is the best fit Keplerian
model.
we set the threshold boundary for a statistically significant
signal to be at the level of 104, or 9.21 in log-Bayesian units,
it is clear that this signal is very significant. Our best fit to
the data yields a non-circular eccentricity, only at the level
of around 0.12, and with a stellar mass commensurate with
that of the Sun, the orbiting planet has a minimum mass
of 1.11 MJ. The phase folded velocities and model fit are
shown in Fig. 2
3.2 HD48265
HD48265 is classed in the Hipparcos catalogue as a G5IV/V
star, however we find a ∆MV of 1.9, therefore this star is
a subgiant. Again, it is extremely metal-rich, towards the
top end of the metallicity scale, since it's [Fe/H] is found to
be +0.4 dex. A planetary companion to this star was previ-
ously published in Minniti et al. (2009) and an update to the
planet's orbital parameters was published in Jenkins et al.
(2009). Here we report our latest orbital solution for this
system, including more HARPS velocities and with the ad-
dition of CORALIE and MIKE data.
We found a signal with a period of 780 days and with a
semi-amplitude of 28 m s−1(Fig. 3) in the radial velocities of
HD48265. The log-Bayesian evidence for this signal is 103.1,
well above the significance threshold. Although there is a
trough in the MMSE close to this period when compared
to the surrounding parameter space, it is significantly lower
than the noise floor at shorter periods, in particular around
a period of 20 days (deepest trough), therefore the MMSE
periodogram can not be used to confirm the existence of this
signal. We studied this 20 day trough to look for evidence
of a signal here but nothing was clear. In fact, this trough
was found to be dependent on the noise properties of the
data, since when the jitter parameter was considered in the
periodogram analysis the evidence pointed towards a period
of ∼40 days, close to twice the period, indicating it is an
artifact of the interference pattern from the window function
beating with the real signal in the data.
The evolved nature of the star ensures that it is an in-
active and slowly rotating star and hence it is unlikely that
stellar activity is the source of these signals. The minimum
mass of the planet is found to be 1.47 MJ and a non-zero
eccentricity was found for the signal. We note that a signifi-
cant linear trend was also found that possibly indicates more
companions that await discovery in this system, particularly
at longer orbital periods. The method also indicated that a
shorter period signal may be present, with a period close to
60 days, yet the current data does not allow us to confirm
this as a genuine second planet since the signal disappears
when we introduce the linear activity correlation parame-
ters in the analysis. Also, this signal could be an alias of the
detected planetary signal.
The planetary parameters we find here are generally
in good agreement with those published in Minniti et al.
(2009) and Jenkins et al. (2009), however the period we
quote is significantly larger than that found by Jenkins et al.
by ∼80 days. The differences in the periods are attributed
to the lower number of high precision HARPS data in that
work, which was causing the fitting algorithm to weight
heavily towards those data points, even though they were
much fewer compared to the MIKE data. This had the effect
of significantly increasing the precision of the fit compared
to Minniti et al., but in the presence of a linear trend, it also
skewed the orbital period to lower values. All other param-
eters are in excellent agreement within the uncertainties.
3.3 HD68402
The star HD68402 is classed as G5IV/V, though since we
find a ∆MV of around 0.1 magnitudes, we believe the
star to be a dwarf. The metallicity of HD68402 is found
to be +0.29 dex. A signal with a semi-amplitude of over
50 m s−1was found at a period (∼1100 days) similar to that
in the HD9174 data. The MMSE shows two long period
troughs for this data, with the detected signal found to be
the second trough behind a slightly stronger trough close to
3000 days. The signal was found to have an eccentricity of
0.03, but to within the limits it can be considered as circular.
Given the mass of HD68402 is 1.12 M⊙ , we compute a
minimum mass of just over 3 MJ for the planet. The phase
folded velocities for the planetary signal are shown in Fig. 4.
Since the log-Bayesian evidence is 49.4, strongly confirming
the existence of a signal, and it was also found in the MMSE
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
7
Figure 5. Phase folded CORALIE (blue) and HARPS (red) ve-
locities for HD72892. The solid curve is the best fit Keplerian
model.
Figure 6. Phase folded CORALIE (blue) and HARPS (red) ve-
locities for HD128356. The solid curve is the best fit Keplerian
model.
analysis, and by a manual fitting approach, the solution is
significantly well constrained, even though there is a small
gap in the phase folded curve.
3.4 HD72892
This G5V star is located at a distance of 73 pc from the
Sun and has a metallicity of +0.25 dex. The star is also
very inactive (logR′
HK = -5.02) and a slow rotator (vsini =
2.5 km s−1) representing an excellent target to search for
planets. We have detected a signal with a period of nearly
40 days and semi-amplitude of 320 m s−1. The signal has a
log Bayesian evidence value of 903.0, confirming the signal
at very high significance. Given the S/N ratio between the
signal amplitude and the HARPS and CORALIE uncertain-
ties, we fully expected a large evidence ratio to confirm the
nature of the signal. The signal is also clearly apparent in
the MMSE periodogram as the strongest trough, adding to
its reality.
We find a mass for the star of 1.02 M⊙ , leading to a
planetary minimum mass of 5.5 MJ. This super-Jupiter also
has appreciable eccentricity, at the level of 0.4, and the final
Keplerian model to the phase folded velocities are shown
in Fig. 5. We note that the solution gives rise to a transit
probability for this star of 1.6%, a relatively high likelihood
of transit for such a planet-star separation.
3.5 HD128356
HD128356 is the coolest star we have included in the CHEPS
sample and is classed as a main sequence star by Hipparcos
(K3V). We find the star to have a [Fe/H] of almost +0.2 dex,
and with a ∆MV of 0.55, it may be a subgiant star, or at
least in the process of evolving onto the subgiant branch.
Using our Bayesian approach we detected a signal
with a semi-amplitude of 37 m s−1and period approaching
300 days that had a log-Bayesian evidence ratio of 144.5.
The star is found to be a very slow rotator, having a vsini of
only 1.3±0.2 km s−1, and with a low chromospheric activity
of logR′
HK of -4.8. We note that this activity level is signifi-
cantly lower than the one reported in Jenkins et al. (2011b),
due to an updated B − V colour used here. The colour used
MNRAS 000, 1–20 (2002)
in Jenkins et al. was drawn from the Hipparcos catalogue
(B − V = 0.685), yet the Tycho-2 catalogue magnitudes
(Høg et al. 2000) give a colour > 1, agreeing with what is
expected for a mid-K star. Even if the star was moderately
active, we would still expect the jitter to be low since mid-K
type stars are not as Doppler−noisy as earlier type stars for
a given activity level (e.g. Isaacson & Fischer 2010). This,
combined with the very slow rotation of the star, indicates
that it is likely a signal with a peak-to-peak amplitude of
nearly 80 m s−1is not caused by modulated activity effects,
especially with a period close to 300 days and eccentricity
of 0.8.
The signal detected in our MCMC analysis is also
clearly apparent in the MMSE, being the deepest trough,
despite the high eccentricity. However, an additional trough
at a much longer period is also approaching a similar level
of significance. There is also no apparent correlations with
the activity indicators, as discussed in the next section, nor
are there any detected periods in the activity measurements.
Given the multi-method signal detection and lack of activ-
ity correlations, we can confidently conclude that the signal
we detected has a Doppler origin, and since the mass of
HD128356 was found to be around 0.65 M⊙ , the measured
minimum mass for the planet is 0.9 MJ.
3.6 HD143361
The star HD143361 was previously shown to have a
planet with a period of 1057 days (Minniti et al. 2009;
Jenkins et al. 2009) and we have been conducting further
reconnaissance to search for additional companions and to
better constrain the orbital characteristics of the previously
detected planet. The host star is a chromospherically quiet
(logR′
HK = -5.12 dex) and metal-rich ([Fe/H] = +0.22 dex)
G6V star, located at a distance of 66 pc.
Our Bayesian search found a signal with a period of
1046 days with a Bayesian evidence of 491.9, relating to a
planet orbiting the star with a minimum mass of 3.5 MJ
(Fig. 7). From the MMSE periodogram the signal is clearly
detected, being one of the most significant periodogram de-
tection's in our sample. The final parameters are in good
agreement with those published in Minniti et al. and Jenk-
8
J.S. Jenkins et al.
Figure 7. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for HD143361. The solid curve is the
best fit Keplerian model.
ins et al. The orbital period found here is lower by 40 days
(∼4%) compared to that published in Minniti et al. but
only lower by 11 days (∼1%) to that published in Jenkins et
al., and both are in agreement within the quoted uncertain-
ties, which are a factor of 28.1 and 6.3 lower here than in
those previous two works, respectively. Although in agree-
ment within the quoted uncertainties, our semi-amplitude
is higher than those published in the previous two works,
by 9.1 and 7.1 m s−1, respectively. No strong evidence for a
second companion was found in this system with the current
data set.
3.7 HD147873
The star HD147873 is the earliest type star in this sample of
planet-hosts and is reported as a G1V star in the Hipparcos
catalogue. Given its distance of 105 pc, the star is a little
brighter than 8 magnitudes in V . The star also appears to
have a solar metallicity ([Fe/H] = -0.03 dex), is extremely
inactive (logR′
HK = -5.27 dex), and rotates at the level of
nearly 6 km s−1. We find a Y2 evolutionary track mass for
HD147873 of 1.38 M⊙ .
The Bayesian search for signals in the Doppler data
for this star detected two strong periodic signals with semi-
amplitudes of 170 and 50 m s−1for HD147873b and c re-
spectively. The periods of the signals were found to be at
117 days for the stronger signal and 492 days for the weaker
of the two signals. The log-Bayesian evidences we found
for these signals were 1131.5 and 145.2. The MMSE peri-
odogram also detected both these signals rather easily; the
second becoming detectable in the residuals of the data once
the first signal was removed, as shown in Appendix B1. We
find planetary minimum masses of 5.1 and 2.3 MJ for the
short and longer period planets, respectively. Both Keplerian
fits to the data are shown in the upper and middle plots in
Fig. 8. The inner planet is also only one of two in this sam-
ple that has a transit probability of over 2%, a value that
encourages the search for transits from intermediate period
planets.
Given we have discovered two giant planets with a
semimajor axis difference of only 0.84 AU between them,
we decided to test if the system architecture was dynami-
Figure 8. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for HD147873 for the short (upper panel)
and long (lower panel) period signals. The solid curves represent
the best fit Keplerian models to the data.
cally stable. We ran Gragg-Burlich-Stoer integrations in the
Systemic Console (Meschiari et al. 2009) over a period of
10 Myrs to study the evolution of the orbits of both planets.
We find the eccentricity of the orbits librate with a period
of around 12000 years but the system itself remains dynam-
ically stable across this timespan. Systems with multiple gi-
ant planets are interesting laboratories for dynamical stud-
ies and this system may warrant further detailed dynamical
study, especially if more massive companions are discovered
with the addition of more data.
3.8 HD152079
HD152079 is classed as a G6 main sequence star in the
Hipparcos catalogue and our previous work found it to be
inactive (logR′
HK=-4.99 dex) and metal-rich (+0.16 dex in
[Fe/H]), which may explain the 0.5 magnitude ∆MV . We
found the mass of the star to be 1.1 M⊙ . This is also one of
the stars in this sample with a previously announced planet
candidate detected in orbit (Arriagada et al. 2010).
A signal with a period of 2900 days and semi-amplitude
of 31 m s−1was detected in the Doppler data of HD152079.
The log-Bayesian evidence ratio was found to be 99.1, highly
significant, and the signal was found to have an eccentric-
ity over 0.5. It is likely for this data set that the moder-
ate eccentricity of the signal is hampering its detection in
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
9
Figure 9. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for HD152079. The solid curve is the
best fit Keplerian model.
Figure 10. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for HD154672. The solid curve is the best
fit Keplerian model.
the MMSE periodogram. In addition, there is the presence
of a linear trend in the data that indicates there is a long
period secondary companion to this star, and since linear
trends are not considered in the MMSE model, the interfer-
ence here could also be confusing the algorithm. Yet there
is a fairly strong trough showing at a period of ∼1400 days,
which is close to half the Bayesian detected signal, and could
be related to the Doppler signal, or an additional compan-
ion that is at a 2:1 resonance site, which could also explain
the eccentric shape of the one planet signal (Marcy et al.
2001; Anglada-Escud´e et al. 2010). In any case, we found
the minimum mass of HD152079b to be 2.2 MJ and the Ke-
plerian model fit is shown in Fig. 9. These values are in
good agreement with those presented in Arriagada et al.,
except the precision quoted here is much higher. For in-
stance, the period of 2097±930 days quoted in their work
has been constrained to 2899±52 days here, a factor of 18
increase in precision and pushing the planet's orbit upwards
by nearly 900 days. This precision increase is also mirrored
directly in the semi-amplitude precision, lowering it from
58±18 m s−1to 31.3±1.1 m s−1.
3.9 HD154672
This star has a Hipparcos classification of G3IV, confirmed
by our measurement of 0.94 magnitudes from the main
sequence. Part of the elevation from the main sequence
can also be explained by the metallicity enrichment of
0.11 dex. The star is also a slow rotator, having a vsini of
2.2 km s−1and was found to be very chromospherically inac-
tive (logR′
HK= -5.12 dex). The position on the HR-diagram
gives rise to a mass of 1.08 M⊙ .
A signal with a period of 164 days and semi-amplitude
of 176 m s−1was found in the Doppler timeseries of
HD154672, with a log-Bayesian evidence ratio of 1709.2,
the most significant signal in this data set. The signal is
also clearly apparent in the MMSE periodogram, the period
trough being significantly stronger than any other periods
across the parameter space. The eccentricity of the signal
was found to be 0.6, giving rise to a planet with a minimum
mass of nearly 5 MJ. The values we find are in good agree-
ment with the values previously published for this planet
MNRAS 000, 1–20 (2002)
Figure 11. Phase folded CORALIE (blue) and HARPS (red)
velocities for HD165155. The solid curve is the best fit Keplerian
model.
in L´opez-Morales et al. (2008), with the period agreeing to
within one hour and the minimum mass being slightly lower
here by only 0.23 MJ, but well within the 1σ uncertainties.
Given the inclusion of higher quality data in this analysis,
we find the jitter for this star to be 2 m s−1, a factor two
lower than that quoted in Lopez-Morales et al., demonstrat-
ing that a significant fraction of their jitter was instrumental
noise. The model fit is shown in Fig. 10.
Our search for additional planets in the combined data
sets did not yield any positive results, therefore no firm
evidence exists for any additional companions in this sys-
tem. If the eccentricity from this planet is genuine and not
due to the super-position of mixed signals from other giant
planets in resonant orbits (see Anglada-Escud´e et al. 2010;
Wittenmyer et al. 2012), then the transit probability for this
object is found to be the highest in the current sample of
intermediate and long period planets, at 2.5%.
3.10 HD165155
The Hipparcos catalogue classifies HD165155 as a G8 main
sequence star, however with a elevation above the main se-
quence of 1.4, this star can be considered as a subgiant. The
10
J.S. Jenkins et al.
Figure 12. Phase folded CORALIE (blue), HARPS (red), and
MIKE (green) velocities for HD224538. The solid curve is the best
fit Keplerian model.
star is located at a distance of 65 pc, and from spectroscopy
we have found a logR′
HK activity index of -5.18 dex, a rota-
tional velocity of 1.5 km s−1, and a [Fe/H] metallicity index
of 0.09 dex. Comparison to Y2 evolutionary models yield a
mass for the star of 1.02 M⊙ .
A signal has been detected in the radial velocity data for
HD165155 with a period of 435 days and a semi-amplitude
of 76 m s−1. The log-Bayesian evidence for the signal was
found to be 168.9, securely above the significance threshold.
The eccentricity was found to be 0.20 and therefore the fi-
nal minimum mass of the companion is calculated as 2.9 MJ
(Fig. 11). A two-planet model search produced statistically
significant evidence for a second signal in the data, however
given the limited number of measurements we could not con-
firm a unique secondary signal at this time. No statistically
significant troughs were detected in the MMSE periodogram
for this star, which may be due to the presence of a sec-
ondary signal that is interfering with the primary signal.
Indeed, the inclusion of a strong linear trend was necessary
to constrain this signal, and since linear trends are not in-
cluded in the MMSE modeling approach, this trend is likely
the reason why the MMSE approach failed to detect this
signal.
3.11 HD224538
The main Hipparcos catalogue lists HD224538 as a F9 dwarf
or subgiant located at a distance of 78 pc. With a calculated
∆MV of 0.63 and a high overabundance of metals in the star
([Fe/H]=+0.27 dex), the possibility remains that this star
is either on the main sequence or crossing into the subgiant
branch. The star is both a slow rotator (vsini = 3.9 km s−1)
and chromospherically inactive (logR′
HK= -4.99 dex). From
comparisons to Y2 isomass tracks on a HR-diagram we found
a mass of 1.34 M⊙ for HD224538.
Our Bayesian algorithm found a signal with a period of
1189 days, a semi-amplitude of 107 m s−1, and an eccentric-
ity of 0.46, shown in Fig. 12. The log-Bayesian evidence ra-
tio for the signal was found to be highly significant at 391.0.
The MMSE periodogram also clearly detected this signal.
Therefore a planet with a minimum mass of 6.0 MJ is found
to be orbiting this star. This is reminiscent of the gas giant
Figure 13. The four plots from top to bottom show the linear
correlations between the radial velocities and the bisector span
velocities, the S-indices, the Hα indices, and the He I indices for
HD128356, respectively, where CORALIE data is represented by
open rings and HARPS data by filled circles. The solid lines are
the best fit linear trends to the data.
planet 14 Her b that has a broadly similar mass, period, and
eccentricity (Butler et al. 2003) and such planets appear to
be rare. Even though 14 Her is an early K-dwarf star it does
have a super solar metallicity (+0.43±0.08 dex) similar to
HD224538, likely necessary to form such high-mass planets.
No additional statistically significant signals were found in
the current data set.
3.12 Line Modulation Tests
Although these stars are very inactive and slowly rotating
and the Doppler signals we have detected are generally very
large compared to the uncertainties (most are significantly
larger than 20 m s−1), it is useful to rule out line modulations
that could originate from stellar activity as the source of
the variations. The activity parameters employed are the
calciumiiHK line doublet, the bisector span (BIS), the CCF
FWHM, the Hα line, and the He I D3 line. These indices
were selected since they have previously been shown to be
good tracers of stellar magnetic activity, and/or spectral line
modulations (e.g. Queloz et al. 2001; Robertson et al. 2014;
Santos et al. 2010). In Fig. 13 we show four of the tests we
have carried out to rule out these modulations as the source
of the detected radial velocity shifts for the star HD128356,
originally believed to be the most active star in the sample
due to the erroneous B − V colour. We note that we do not
show the CCF FWHM test for this star since there is large
variations with a few outliers, but no correlation exists.
In the upper plot of Fig. 13 we show how the BIS values
vary as a function of the radial velocity datasets. The BIS
values for HARPS were taken from the HARPS-DRS and
measured following the method explained in Queloz et al.
(2001). The CORALIE BIS values were calculated using a
similar procedure. No significant correlation between the ra-
dial velocities and the BIS measurements are found and we
highlight this by showing the best fit linear trend to the
MNRAS 000, 1–20 (2002)
data. The unweighted Pearson rank correlation coefficient
has a value of 0.23, signifying a weak correlation, however
when the correlation is weighted by the measurement uncer-
tainties on the radial velocity and BIS values, the coefficient
drops to 0.11, or no evidence at all for any correlation.
We also searched for a correlation between the chro-
mospheric activity S-indices and the velocities as a second
useful discriminant that activity is not the source of the
observed variations. The measurement of these S-indices
for HARPS was briefly discussed in Jenkins et al. (2013b)
and therefore here we only discuss the CORALIE activ-
ity measurement method in Appendix D. In any case the
method for both is similar, except for HARPS we use the ex-
tracted 1D order-merged spectrum, whereas for CORALIE
spectra we perform the calculations using the extracted 2D
order-per-order spectrum, similar to the method discussed
in Jenkins et al. (2006).
The second plot in Fig. 13 shows these chromospheric
activity S-indices as a function of the radial velocity mea-
surements and no apparent correlation is found. The best
unweighted linear fit is shown by the solid line and confirms
the lack of any correlation between the two parameters. The
correlation coefficient also confirms this since an unweighted
r coefficient of -0.09, similar to the weighted BIS, is not sta-
tistically significant, dropping even lower when considering
the weights.
The lower two plots in the figure show the linear corre-
lations against the measured Hα and He I D3 activity indi-
cators, respectively. These indices were calculated following
the methods discussed in Santos et al. (2010). For both of
these indices, no significant correlation is found when com-
bining the CORALIE and HARPS data. Some moderate
correlation between the He I index and the velocities is seen,
with an unweighted r correlation coefficient of -0.64 for the
HARPS only measurements, decreasing to -0.39 when the
CORALIE measurements are added. Judging by the lower
panel in the figure, no striking correlation is apparent, given
what would be expected for this level of correlation, and
once the measurement uncertainties are included to weight
the correlation coefficient, the value drops significantly to be
in agreement with zero correlation. In fact, we can see that
the majority of the data are uncorrelated, from radial veloc-
ities between -25 - +55 m s−1, with only a few offset data
points clustered around -40 m s−1driving the correlation.
As an aside,
relationships
if we apply the
in
Saar & Donahue (1997) and Hatzes (2002) to calculate the
spot coverage expected for a star with the rotational period
of HD128356, in order to produce a radial velocity ampli-
tude in agreement with that observed here, then ∼5% of
disk spot coverage is required, which would likely exhibit as
photometric variations that are not observed (see below).
Although the activity indicators for the other stars re-
ported in this work show no evidence for any strong linear
correlations, measured by the Pearson Rank correlation co-
efficient, against the radial velocities, we report the moder-
ately correlated data sets (0.5 > r > 0.75). For HD48265
the HARPS BIS values correlate with the radial velocities
with a r value of 0.59±0.25, indicating some moderate corre-
lation between the two parameters. We also note that both
the Hα and He I indices have values of 0.47 and -0.41, re-
spectively, yet there are large parts of parameter space that
are under-sampled by including only the HARPS data alone.
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
11
When adding in the CORALIE measurements we find these
values decrease to 0.22, 0.12, and -0.33 for the three quanti-
ties respectively, with uncertainties of ±0.16, indicating that
these correlations are not the source of the velocity signal
for this star.
The star HD68402 shows HARPS velocity correlations
with the BIS and He I indices with r values of 0.79 and
0.52, respectively. We note that there are only five HARPS
data points for this star so no result here can be deemed
significant. Furthermore, once the CORALIE values are in-
cluded we find values of 0.01 and 0.12 for these parameters
respectively. When including the CORALIE measurements
we find a moderate correlation appears between the veloc-
ities and the CCF FWHM measurements (r = 0.52±0.23).
Again there appears no significant correlations in the anal-
ysis.
Another star with a limited number of HARPS spectra
(eight measurements) that give rise to an apparent moderate
correlation between the radial velocities and activity indica-
tors is HD72892. The HARPS CCF FWHM and He I mea-
surements have Pearson rank correlation coefficients of 0.58
and 0.73, respectively, with uncertainties of ±0.38. When
the CORALIE measurements are added to the HARPS data
we find this correlation becomes insignificant, with values of
only -0.47 and 0.10, respectively, and uncertainties of ±0.19.
However, the Hα indices now exhibit a moderate correlation
with the velocities (r = 0.57). The signal for this star has
an amplitude of nearly 320 m s−1and a period of ∼40 days.
Although 40 days is a plausible rotational period for this
type of very inactive G5 dwarf star, the fact that it is so
inactive indicates that such a large signal would be difficult
to produce through spot rotation. In fact, if we calculate the
spot coverage as above, then 20-25% of disk spot coverage is
required, which can be ruled out based on the photometric
stability and the low logR′
HK measurement.
For HD152079 the HARPS Hα index correlates with
the velocities with a correlation coefficient r of 0.66±0.24,
although none of the the other indicators show any evidence
for correlations, and when adding the CORALIE data, the
correlation coefficient decreases to a value of only 0.47±0.17.
We note that the variations in the measurements are only
changing at the few ×10−3 level, from 0.203 to 0.208 in our
HARPS Hα index.
The star HD165155 shows a correlation between the
velocities and the HARPS CCF FWHM with a value of -
0.66±0.23, although with a large spread. Addition of the
CORALIE CCF FWHM decreases the value to 0.12±0.17,
rendering this result insignificant.
HD224538 shows moderate correlation between the ve-
locities and the Hα indices at the level of 0.69±0.22 in
the HARPS data, and at the level of 0.27±0.15 when the
CORALIE values are included. Again the variation is only
at the few ×10−3 level, which is likely to be insignificant.
The periodogram analysis for each of the five activity
indicators did not reveal any significant periods that could
explain the detected signals in any of the stars considered
here (see Appendix E1), however a few features do appear.
For the star HD72892 there are emerging peaks in the peri-
odograms of the BIS, FWHM, S and He I indices at periods
between 11-13 days, and although this region is distinct from
the detected planetary signal, these may be linked to the ro-
tational period of the star, however there is a peak in the
12
J.S. Jenkins et al.
window function at 22.5 days that could be giving rise to a
peak at the first harmonic in these indices. For HD128356,
the moderately active star, both the BIS and FWHM time-
series show peaks that agree with a period around 1250 days
that has no counterpart in the window function, meaning
this could be a magnetic cycle, but it is far from the de-
tected Doppler signal period. HD147873 does show a unique
peak close to the lower period signal in the radial velocities
in the He I indices with a period of around 120 days, yet
the peak is not significant. The Hα indices show an emerg-
ing peak with a period of ∼70 days as the strongest signal
for HD165155 that could point to the rotation period for
this star but there is a window function peak emerging at
63.5 days, that is likely unrelated to this peak, but is wor-
thy of note. In any case, neither of these are related to the
detected Doppler signal in the radial velocities. Finally, the
star HD224538 shows evidence for two peaks in the peri-
odograms of the BIS and Hα indices that are in good agree-
ment with signals with periods of 20 days, which again could
be a good candidate for the rotational period for this star,
however the fifth strongest peak in the window function is
found to be at 19 days, meaning there is a non-insignificant
probability that this peak is being boosted by the sampling.
3.13 Photometric Analysis
We decided to photometrically search for a secure rotational
period for these stars by employing frequency analyses of the
V -band All Sky Automated Survey (ASAS; Pojmanski 1997)
photometric data. We have previously shown that such anal-
yses can shed light on the rotational periods of planet-host
stars and/or short period and long period magnetic cycles
(e.g. Anglada-Escud´e et al. 2013; Jenkins & Tuomi 2014).
We tend to focus on the best quality data, ASAS grade A or
B using the smallest ASAS apertures that are best for point
sources, and typical baselines cover ∼9 years at a sampling
cadence of ∼3 days. Out of all the 11 planet-hosts considered
in this work, six show evidence for a significant rotational
period or long period magnetic cycle in the photometry, and
these are summarised below, with particular focus paid to
HD165155.
The six stars showing evidence for a photometric signal
in the ASAS timeseries are HD68402, HD147873, HD152079,
HD154672, HD165155, and HD224538, and the Lomb-
Scargle periodograms for these six stars are shown in Ap-
pendix F1. We do not show the periodograms for the re-
maining six stars since they exhibit no significant frequency
peaks. The periodogram for HD68402 shows two strong
peaks emerging at periods of 312 days and 2000 days. Nei-
ther of these signals reside at periods close to the detected
Doppler signal in the radial velocities, however their strength
suggests there may be some long term magnetic cycle at play
within this star.
The stars HD147873, HD152079, HD154672, and
HD224538 all show evidence for long period modulated spot
activity, with periods at the extremities of the data time-
series and periodogram sampling ∼5000–10000 days. These
could be real long term spot cycles or they could be sam-
pling features to due to the limited baselines of the data sets.
However, none of these features appear to coincide with the
detected radial velocity signal periods, or harmonics there-
of. HD154672 has a velocity signal detected at just over
160 days, far from any long period magnetic cycle, whereas
the signal in the HD224538 velocities has a period of a few
thousand days, which could agree with any potential pho-
tometric signal in the ASAS data that is not due to the
limited data baseline. However, our analysis never indicated
any correlations were evident between the radial velocities
and the activity indicators for this star, and its inactive na-
ture, along with the strength of the radial velocity signal
(K = 110 m s−1), would make an activity origin unlikely
for a star of this type. The evidence from these analyses
points to the origin of the detected signals as being due the
gravitational influence of orbiting planets.
The star HD152079 has a detected signal in the veloci-
ties with a period of a few thousand days, which is approach-
ing the regime where a long period magnetic cycle could be
present due to the ASAS photometric periodogram, yet the
amplitude of this signal is a little over 30 m s−1for this very
inactive star. The structure of the long period signals in the
power spectrum of these stars are very similar, which argues
that the frequencies emerge due to the sampling baseline.
A further secondary peak exists in the photometric peri-
odogram for this star at ∼830 days. Since this is too short
to be associated with the signal in the radial velocities, it is
not the origin of that signal but could be a possible sampled
magnetic cycle. In any case, the nature of the radial velocity
signal is likely Doppler and from an orbiting planetary mass
candidate.
We note that for HD147873 and HD224538, additional
peaks arise in the periodograms at periods that could be
in the range of rotational periods for stars with these types
of rotational velocities and stellar radii, or could relate to
additional magnetic cycles. For HD147873, there is a strong
peak at a period of 29.5 days, very close to the lunar cy-
cle, and since this period was found to arise in other ASAS
timeseries, it is likely this is a sampling alias and not the
star's rotational period. For HD224538, the next strongest
peak is located at 385 days, with again the 29.5 day period
being detected. The 385 day peak is within a small cluster
of peaks that surround the Earth's orbital period, therefore
it is likely this is another sampling alias. Hence, it is unlikely
that we have made a significant detection of the rotational
period for any of these stars, with only tentative detections
of long period magnetic cycles.
Finally, we discuss the photometric analysis for the star
HD165155 independently, since there is an indication of a
peak in the periodogram that is close to the period of the
detected signal in the RV measurements. Given that the sig-
nal is rather strong, it is unlikely that activity is the source
of the signal at this type of period. We also note that the
orbital separation is too large to produce star-planet inter-
actions that could cause any photometric signal. From the
ASAS periodogram in Appendix F1 we can see that again
a long period signal emerges, but after this signal, there are
two fairly strong peaks with periods of 454 and 344 days.
The 454 day signal is the second strongest after the long
period peak and it closely matches the period of the sig-
nal in the radial velocities at 452 days, indicating the signal
could arise from activity or pulsations. As mentioned above,
no similar periodicities were found in the activity indicators
and since the Hipparcos photometry for this star only con-
sists of 67 measurements, with a scatter in the data of 0.019
magnitudes, no significant periodicity was found in this data
MNRAS 000, 1–20 (2002)
either. We also note that by removing the long period trend
with periods of 10000 days or more removes the signal at
454 days, signifying it is linked to this long period trend
and therefore likely not a true magnetic cycle that could
induce such a radial velocity signal as we observe.
There exists a small possibility that the detected radial
velocity signal is not of Doppler origin and is due to line
asymmetries from stellar activity on this subgiant star, even
though the existence of this photometric period is difficult
to causally connect to the origin of the radial velocity signal
without corroborating periodicities in the spectral activity
indicators. Without garnering more data, and since the de-
tected photometric peak in the periodogram could be an
alias that is associated with a longer period signal or the
window function of the data, we still consider the radial ve-
locity signal as due to an orbiting companion. If, on the
other hand, the signal in the velocities is genuinely of astro-
physical origin, this data set would serve as a warning when
trying to understand the origin of signals in radial velocity
timeseries of subgiant stars, even when the signal amplitude
is relatively large, and there are no correlations or periodic-
ities in the spectral activity indicators. Thorough searches
of existing photometric data should always be performed,
where possible, to help to confirm the reality of proposed
planetary systems.
4 PLANET POPULATION DISTRIBUTIONS
The high metallicity selection bias in our program means we
are generally targeting gas giant planets. The working hy-
pothesis being that if such planets are formed through core
accretion processes, then large cores can form quickly due
to the enrichment of the proto-planetary disk, which gives
the planetesimals sufficient time to accrete gaseous material
to reach large masses after they cross the critical core mass
limit of around 10 M⊕(Mizuno 1980).
4.1 Mass Function
The observed mass distribution is a key observational con-
straint for planet formation models, a constraint which has
previously been fit by smooth power law trends with indices
around -1 (e.g. Butler et al. 2006; Lopez & Jenkins 2012). In
Fig. 14 we show the results of applying an exponential func-
tion to the data, which we found to be more suited to the
current distribution of exoplanets that have been detected
over a wide range in stellar mass
f (m) = A × emsin(i) + B
(2)
where f (m) is the model function that we fit to the
data and A and B are the scaling parameter and offset of the
model that are left as free parameters to be found following a
maximum likelihood procedure with the following Gaussian
likelihood function:
L(Θ) = −0.5 × log(2π) −Xi
log(σt,i) − Pi(yi − f (m)i)2
σ2
t,i
(3)
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
σt,i = qσ2
p,i + σ2
e,i
13
(4)
Here L is the likelihood function for parameters Θ, y is
the observed data (mass function histogram points) for all
i histogram points, and σp and σe are the Poisson uncer-
tainties and any excess uncertainty for each of the values,
respectively. This procedure finds the following values for the
modeled parameters 0.89±0.03, 0.030+0.004
−0.003 , and 0.034+0.009
−0.002
for A, B, and σe, respectively. The uncertainties on these
parameters were determined using a Markov Chain Monte
Carlo (MCMC) procedure in Python, employing the emcee
numerical package (Foreman-Mackey et al. 2013). We used
100 walkers and ran chains of 10000 steps in length, with
a 1000 step burn-in, which relates to a final chain length
of 900,000 steps, with a final mean acceptance rate of 49%.
The parameter values we measure are insensitive to small
changes in the bin size used in the histogram, which we set
to be 0.5MJ, a value that allows enough samples in most of
the bins to reflect the smoothly varying distribution.
At the right of Fig. 14 we show the parameter extent
probed by the chains, where we used uniform priors for
the parameters except the excess uncertainty, where we em-
ployed a Jeffries prior where the probability is proportional
to 1/σ. The distribution of the parameters are well confined
to the region around the maximum likelihood value for each,
showing the model we put forward is an acceptable repre-
sentation of the current exoplanet mass function. We note
that the A and B parameters follow Gaussian distributions,
whereas the excess noise parameter is more like a skewed
Gaussian or Poisson distribution, where the lower 1σ credi-
bility limit is found to be close to the maximum likelihood
value of 0.03. In any case, it seems that the mass function
appears to be fairly well described by an exponential func-
tion.
4.2 Mass-Metallicity Functions
As Fig. 15 shows, we tested if there was any metallicity
dependence in the mass function. In order to test this we
split the sample into three metallicity bins, a high metallicity
bin ([Fe/H]>+0.2 dex), an intermediate metallicity bin (-
0.16[Fe/H]<+0.2 dex), and a low metallicity bin ([Fe/H]<-
0.1 dex). These bin sizes allowed a useful number of samples
in each bin to statistically probe the distributions.
Metallicity splitting gives us probabilities (D-statistics)
from two-tailed KS-tests of 8% (0.165) that the high metal-
licity bin and the low metallicity bin are drawn from the
same parent population, and 6% (0.161) that the interme-
diate metallicity planet-hosts and the low metallicity planet
hosts are also drawn from the same population. By com-
bining the high metallicity bin and the low metallicity bin
values and comparing those to the intermediate metallic-
ity bin, the probability is essentially the same, only drop-
ping the D-statistic by 0.01, with a probability of only 6%
that the two populations are statistically similar. To perform
this test we decided to remove the lowest mass planets from
the metal-poor and intermediate-metallicity samples since
Jenkins et al. (2013b) shows that there appears to be a cor-
relation between the mass and metallicity in the low-mass
regime. Therefore, we only consider planets with minimum
masses above 0.0184 MJ as this is the lowest mass planet in
the high metallicity sample, neglecting the exceptional case
14
J.S. Jenkins et al.
Table 2. Orbital mechanics for all planetary systems described in this work.
Parameter
HD9174b
HD48265b
HD68402b
HD72892b
HD128356b
HD143361b
Orbital period P (days)
Velocity amplitude K (m/s)
Eccentricity e
ω (rad)
M0 (rad)
m sin i(MJ )
Semimajor axis a (AU)
γHARPS (m/s)
γCORALIE (m/s)
γMIKE (m/s)
σHARPS (m/s)
σCORALIE (m/s)
σMIKE (m/s)
γ [ms−1year−1 ]
PT
NObs
ln B(k, k − 1)
1179±34
20.8±2.2
0.12±0.05
1.78±0.66
3.5±1.3
1.11±0.14
2.20±0.09
-7.2±1.4
-1.6±3.0
–
1.8±0.6
2.2±1.0
–
–
0.3%
29
42.1
780.3±4.6
27.7±1.2
0.08±0.05
6.0±2.4
4.9±1.4
1.47±0.12
1.81±0.07
-1.5±1.6
-4.3±2.6
-3.5±1.4
6.0±0.6
2.7±1.1
2.8±0.8
–
0.5%
57
103.1
1103±33
54.7±5.3
0.03±0.06
0.3±2.3
6.0±2.2
3.07±0.35
2.18±0.09
-34.2±8.2
-10.6±4.6
–
1.7±0.9
2.0±1.0
–
–
0.2%
20
49.4
39.475±0.004
318.4±4.5
0.423±0.006
6.010±0.014
2.714±0.010
5.45±0.37
0.228±0.008
-37.8±1.7
48.7±3.2
–
2.2±0.7
2.0±1.0
–
–
1.6%
32
903.0
298.2±1.6
36.9±1.2
0.57±0.08
1.47±0.08
3.1±0.7
0.89±0.07
0.87±0.03
-0.1±1.9
9.4±2.7
–
3.9±0.7
2.1±1.0
–
–
0.4%
60
144.5
1046.2±3.2
72.1±1.0
0.193±0.015
4.21±0.06
3.21±0.14
3.48±0.24
1.98±0.07
-1.2±0.8
3.4±2.2
-26.6±1.2
2.3±0.6
1.8±0.9
2.8±0.8
–
0.2%
80
491.9
Parameter
HD147873b
HD147873c
HD152079b
HD154672b
HD165155b
HD224538b
Orbital period P (days)
116.596±0.023
491.54±0.79
Velocity amplitude K (m/s)
Eccentricity e
ω (rad)
M0 (rad)
m sin i(MJ )
Semimajor axis a (AU)
γHARPS (m/s)
γCORALIE (m/s)
γMIKE (m/s)
σHARPS (m/s)
σCORALIE (m/s)
σMIKE (m/s)
γ [ms−1year−1 ]
PT
NObs
ln B(k, k − 1)
171.5±1.2
0.207±0.013
1.40±0.05
1.65±0.07
5.14±0.34
0.522±0.018
59.0±1.2
5.4±2.2
37.6±3.1
2.6±0.7
1.9±0.9
2.4±1.0
2.94±0.68
2.3%
66
1131.5
47.9±1.7
0.23±0.03
0.73±0.20
3.09±0.20
2.30±0.18
1.36±0.05
–
–
–
–
–
–
–
0.7%
–
145.2
2899±52
31.3±1.1
0.52±0.02
5.67±0.06
0.8±0.8
2.18±0.17
3.98±0.15
-37.9±7.0
-44.8±8.5
-13.6±6.8
1.5±0.6
2.0±1.0
2.7±0.8
1.72±0.47
0.1%
50
99.1
163.967±0.009
176.3±0.7
0.600±0.004
4.63±0.01
3.60±0.02
4.73±0.32
0.59±0.02
5.2±0.7
-44.6±2.0
28.2±1.2
2.1±0.5
2.2±1.0
3.6±0.7
–
2.5%
72
1709.2
434.5±2.1
75.8±3.0
0.20±0.03
3.7±0.2
0.9±0.8
2.89±0.23
1.13±0.04
-59.6±18.7
-87.7±20.1
–
5.8±0.6
3.7±1.2
–
4.00±1.19
0.4%
38
168.9
1189.1±5.1
107.0±2.4
0.464±0.022
0.40±0.03
0.3±0.3
5.97±0.42
2.28±0.08
-15.3±1.5
27.0±2.7
55.9±4.3
2.9±0.6
2.0±1.0
5.2±0.7
–
0.2%
50
391.0
The uncertainties on the m sin i and semimajor axis consider the uncertainties on our stellar mass estimate of 10%.
The γ offset is the value after subtracting off the mean of the data set.
The σ terms parameterise the excess noise in our model fits, aka jitter.
The ln B(k,k-1) are generally the 1-planet models (e.g. B(1,0)) except for HD147873 which is a 2-planet model (B(2,1)).
PT is each planet's transit probability.
NObs are the total number of radial velocities per target star.
Exponential Function
0
2
6
4
Msin(i) [MJ]
8
10
Figure 14. Normalised mass function distribution showing an exponential fit to the data for 444 exoplanet candidates (left). The asso-
ciated uncertainties have been calculated assuming Poisson statistics. The right plot shows the parameter space contours and histograms
(aka. a corner plot; Dan Foreman-Mackey et al. (2016). corner.py: corner.py v1.0.2. Zenodo. 10.5281/zenodo.45906) constructed from
the MCMC chains. The contours show the exponential scaling (A), the exponent (α), and the excess noise for 1, 2, and 3σ percentiles
radiating outwards from the point of maximum probability of the distributions. The cross-hairs mark the values determined from the
maximum likelihood best fit. The right-edge plots show the histogrammed distributions collapsed in only the x dimension, where the
mean (solid) and 1σ (dashed) ranges are highlighted.
0.6
0.8
A
1.0
1.2
0.0 2
0.0 4
0.0 6
0.0 8
0.0 3
B
0.0 9
0.0 6
σe
0.1 2
of the planet orbiting Alpha Centauri B (Dumusque et al.
2012) that Hatzes (2013) and Rajpaul et al. (2015) claim
may be attributed to other phenomena like stellar activity
or sampling ghosts.
In order to firm up these statistics we also ran the sam-
ples through the Anderson-Darling (AD) test, which gener-
ally tends to be more sensitive than the standard KS test
since it gives more statistical weight to the tails of the dis-
tribution. From these tests we found p-values of 2% and
5% for the comparison between the high metallicity and
MNRAS 000, 1–20 (2002)
0.6
0.5
0.4
0.3
0.2
0.1
)
m
(
f
d
e
s
i
l
a
m
r
o
N
0.0
0.15
0.10
0.05
0.00
−0.05
−0.10
s
l
a
u
d
i
s
e
R
B
e
σ
0.0 8
0.0 6
0.0 4
0.0 2
0.1 2
0.0 9
0.0 6
0.0 3
New Planetary Systems from the CHEPS
15
(2012), and the overabundance of Jupiter's is not due to the
enhanced formation of such planets as a function of metal-
licity. We did not run the SWEET-Cat samples through the
AD test since the results were shown to be very similar to
the KS tests for the full sample.
4.3 Other Observational Properties
Within the period-mass plane some features can be seen in
the metallicities of exoplanet host stars. An examination of
the left plot in Fig. 16 reveals that there is a broad mix
of metallicities for the gas giant planets and the planets we
publish here are located predominantly in the upper right
quadrant of the plot, with only two having periods below
100 days.
4.3.1 Host Star Metallicities
It appears that the lowest mass planets are found mostly
on short period orbits, due to the inherent sensitivities of
Doppler surveys, and they also appear to orbit metal-poor
stars in general, hence the dominance of the black points to-
wards the bottom left of the left hand panel in Fig. 16. This
result was previously highlighted by Jenkins et al. (2013b),
revealing a 'planet desert' for the most metal-rich stars,
and subsequent confirmation has also been discussed in
Marshall et al. (2014). The nature of this desert could be
explained by core accretion theory whereby the lower den-
sity discs have limited metals to form cores, whereas the high
density discs can readily form high-mass cores that quickly
grow to more massive objects, crossing the critical core mass
limit and becoming gas giants. In fact, the metallic proper-
ties of all planets with periods of less than 100 days appears
to be different when we compare planets more massive or
less massive than 0.1 MJ. In the plot, this is shown by the
significantly higher fraction of red data points above a mass
of 0.1 MJ compared to below that limit.
More directly we can test the reality that the metallic-
ity distribution for planets with periods of 100 days or less
have a different metallicity distribution by again applying
an AD test to the sample of known exoplanets. We chose to
apply the sample to the homogeneous samples that we pre-
viously cross-matched with the SWEET-Cat list. This test
reveals a T-statistic of 8.54 when adjusted for all non-unique
values, revealing a probability of 2×10−4 that these samples
are statistically similar. A Kolmogorov-Smirnov test yields
a similar probability value (4×10−4) with a D-statistic of
0.375. The histograms of both populations are shown in the
top plot of Fig. 17. Although the two histograms appear to
show similar forms, the host stars that contain lower-mass
planets currently has a flatter shape than the host stars con-
taining higher-mass planets in this period space. Although
the sample sizes are small, 48 objects in the low-mass pop-
ulation for example, it does appear that the lowest-mass
planets are drawn from a different metallicity sample when
compared with the most massive planets, within the limits
of the current data set. In the future with many more dis-
coveries of very low-mass planets orbiting the nearest stars
from Doppler surveys, since these represent the most pre-
cise metallicities that can be measured, trends such as those
discussed here can be tested at a higher level of statistical
significance.
Figure 15. The observed mass distribution split into three bins
with different metallicity distributions and a binsize of 2 MJ.
The solid black curve is for the most metal-rich planet-hosts, the
dashed blue curve is for the intermediate metallicity stars, and
the dot-dashed red curve is the distribution for all metal-poor
stars. The metallicity cuts are shown in the key.
low-metallicity samples, and between the intermediate and
low-metallicity samples, respectively. This is in good agree-
ment with the KS test results, indicating that there is a
correlation between mass and metallicity, whereby metal-
rich stars produce many more Jupiter-mass planets com-
pared to super-Jupiters, but metal-poorer stars produce a
higher fraction of super-Jupiters than Jupiters compared to
the metal-rich population. However, the current sample of
host star properties were not drawn from a homogeneous
source and therefore the heterogeneous nature of the data
could be influencing the results.
To try to circumvent this problem, we decided to search
for our sample of exoplanet-hosts in the SWEET-Cat cat-
alogue (Santos et al. 2013). The SWEET-Cat is a project
that plans to eventually contain all exoplanet host star prop-
erties like Teff and metallicity that have been measured using
high resolution spectroscopy in a homogeneous fashion. We
were able to find 93% of our sample in the SWEET-Cat,
but some of these were not measured homogeneously. From
this sample we reran the KS tests and found probabilities
of 9.1% that the high and low metallicity bins are drawn
from the same parent population and 13.8% that the inter-
mediate and low metallicity bins are drawn from the same
distribution.
A further step that was taken was to remove even more
information but improve the homogeneity of the sample. We
selected only those stars with a homogeneous flag of 1 in
the SWEET-Cat, which means that the properties of these
stars were measured using the same general methodology.
This selection resulted in a 20% loss of information but still
contained a total sample size of 358 planet-hosts, however
the low metallicity bin only contained 64 stars, whereas the
high and intermediate bins have sample sizes of 131 and 163
objects, respectively. The KS test probabilities are now sig-
nificantly lower than the full sample, having values of 38.0%
and 52.0% that the high and intermediate mass functions
are statistically similar to that of the low metallicity bin.
These tests likely show that currently there are no statisti-
cally significant correlations between planetary mass and the
metallicity of their host stars, as claimed by Mortier et al.
MNRAS 000, 1–20 (2002)
16
J.S. Jenkins et al.
10.0
0.1
Dataset
Lit
New
[Fe/H]
0.50
0.25
0.00
−0.25
−0.50
−0.75
10.0
0.1
Eccentricity
1.00
0.75
0.50
0.25
0.00
Dataset
Lit
New
]
s
e
s
s
a
M
r
e
t
i
p
u
J
[
i
i
n
s
M
]
s
e
s
s
a
M
r
e
t
i
p
u
J
[
i
i
n
s
M
10
Period [days]
1000
10
Period [days]
1000
Figure 16. Distribution of exoplanet metallicities (left) and eccentricities (right) within the period-mass plane. The small circles are
the literature values and the large circles are our targets from this work. The colour scale to highlight the differing metallicities and
eccentricities are shown at the right of both plots.
to 100 days and those beyond 100 days. The binned his-
togram for both samples is shown in the lower plot of Fig. 17,
and although the distributions appear less discrepant than
the mass cut in the top plot, there is an indication of a
functional change in the metal-poor regime. The AD test
of the metallicities from these samples returns a T-statistic
of 7.62, leading to a probability of 4×10−4 that the sam-
ples are similar. This time the Kolmogorov-Smirnov test
reveals a slightly smaller probability of the null hypothe-
sis, returning a value of 8×10−3 and a D-statistic of 0.223.
Therefore we find that, in general, short period giant plan-
ets have higher metallicities than those at longer periods,
with a mean value of [Fe/H] of 0.16 dex for the sub-100 day
planets and a value of 0.06 dex for the giant planets with or-
bital periods longer than 100 days, as suggested by Sozzetti
(2004) and Pinotti et al. (2005).
Mordasini et al. (2012) constructed global population
synthesis models of forming planets in a range of disc envi-
ronments to search for expected correlations between plane-
tary orbital parameters and bulk compositions against disc
properties. They found that planets tend to migrate more in
low-metallicity discs compared to more metal-rich discs be-
cause the cores that form in the low-metallicity environment
need to migrate more to undergo enough collisions to grow
to the critical mass limit and transition from Type I migra-
tion to the slower Type II migration. They suggest no clear
correlation between semimajor axis, or orbital period, exists
because the planets in low-metallicity discs also form further
from the central star than in the high-metallicity discs, and
so the increased efficiency of migration in the low-metallicity
environment is compensated by the increased distance the
planets need to travel inward towards the star.
These modeling efforts tend to be at odds with the find-
ings we have made unless certain conditions apply. If giant
planets in metal-poor discs migrate more then we would ex-
pect to see the opposite result, unless the planets start their
MNRAS 000, 1–20 (2002)
Figure 17. The binned histogram of sub-100 day period exo-
planet host star metallicities detected by radial velocity programs
is shown in the top plot. The solid histogram represents the gas
giant planets, minimum masses above 0.1 MJ, and the dashed
histogram is for host stars with planets below this threshold. The
lower plot shows the binned histogram of gas giant planets (mini-
mum masses above 0.1 MJ) for orbital periods of 100 days or less
(solid histogram) and those with periods above 100 days (dashed
histogram). All histograms have been normalised to the peak of
the distribution to highlight their differences.
Further to this, the high-mass planet sample may in-
dicate there is a non-uniform metallicity distribution as a
function of period. To test this we split the high-mass planet
sample into two bins with orbital periods less than or equal
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
●
journeys very far out in the disc before arriving at their
current locations. In addition, another scenario could be
that the low-metallicity discs are dispersed faster than high-
metallicity discs through photo-evaporation (Yasui et al.
2009; Ercolano & Clarke 2010). This effect is thought to
be due to the lower optical depth of the disc allowing the
UV and X-ray flux to pass deeper into the disc, dispersing
the inner regions faster, meaning there is no remaining gas
and dust for the planet to interact with, essentially halt-
ing its migration earlier when compared to a planet migrat-
ing through a metal-rich disc. Finally, in the Mordasini et
al. model, they predominantly consider mostly inward mi-
gration of cores, however recent work has shown that disc
structure is important in defining the dominant torques that
drive planet migration and in the inner discs that are heated
by the intensity of the young star's radiation field, corota-
tion torques dominate over the differential Linblad torques,
leading to outward migration of the cores (Kretke & Lin
2012). Further to this, random walk motion too can be im-
portant for migrating low-mass cores (Nelson & Papaloizou
2004; Laughlin et al. 2004; Nelson 2005) and dead-zones in
the disk can subsequently halt the migration of forming low-
mass cores (Balmforth & Korycansky 2001; Li et al. 2009;
Yu et al. 2010). All of these processes could lead to the
preservation of a period-metallicity relationship that favours
short period planets predominantly being found orbiting
more metal-rich stars and longer-period planets being found
in more metal-poor environments.
4.4 Orbital Eccentricities
In the right plot in Fig. 16 we show the same period-
minimum mass plane, yet this time the colour scaling high-
lights the eccentricity distribution. We can see that the ma-
jority of the short period planets (P 610 days) are generally
found to have circular orbits, a fact that can be attributed to
the planets tidal interactions with the host star that tends
to circularised their orbits. We also see that the majority of
the low-mass planets are found on circular orbits too (black
points), in comparison to the high-mass planets where a sig-
nificant fraction of them have moderate-to-high eccentric-
ities (red points). Note that there is also a selection bias
towards the detection of higher eccentricities that depends
on the quantity of radial velocity data points that describes
a given signal (O'Toole et al. 2009), whilst high eccentricity
also elevates the amplitude of a given Doppler signal, which
can sometimes make them easier to detect.
If we again split the planets up into two mass bins,
where the low-mass planets have minimum masses of
60.1 MJ and the high mass bin comprises all planets with
minimum masses above this limit, the eccentricity means
and standard deviations of the two populations are 0.13 and
0.12 for the low-mass planet population and 0.25 and 0.21
for the high-mass planet sample. Taken at face value, the in-
creased standard deviation for the higher mass sample tells
us the spread in eccentricities in this mass regime is higher
than for the lower mass planets. There is a strong bias here
where the low-mass planet sample has a significantly lower
mean orbital period, with a much higher fraction of planets
orbiting close enough to the star to be quickly circularised
through tidal dissipation of the orbits. Furthermore, there is
a tendency to fix the eccentricity to zero when performing
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
17
Keplerian fits to radial velocity data, in order to remove this
additional degree of freedom and the degeneracy with other
parameters being fit at the same time.
5 SUMMARY
We have used the CORALIE, HARPS and MIKE spectro-
graphs to discover eight new giant planets orbiting seven
super metal-rich stars, and a star of solar metallicity, along
with updated orbits for four previously published plan-
ets. We include radial velocity data prior- and post-2014
CORALIE upgrade and our Bayesian updating method re-
turned a systematic offset of 19.2±4.8 m s−1between the
two velocity sets for our stars. The new planets cover a wide
area of the giant planet parameter space, having a range of
masses, periods, and eccentricities, including a double planet
system that was found orbiting the most massive star in our
list, and a 14 Herculis b analogue that has a minimum mass
of ∼5.5MJ, an orbital period of nearly 1200 days, and signif-
icant eccentricity (e=0.46), adding another member to the
sub-population of massive eccentric planets orbiting super
metal-rich stars.
We introduced our method for measuring the chromo-
spheric S-index that is a measure of the magnetic activity
of Sun-like stars using CORALIE spectra. These activities,
along with bisector measurements, CCF FWHM's, Hα in-
dices, HeI indices, and Hipparcos and ASAS photometry,
were used to rule out the origin of the planetary Doppler
signal as being due to line modulations from rotationally
influenced star spot migration or other activity phenomena
like chromospheric plage or stellar pulsations.
We show that the mass function for planets is well de-
scribed by an exponential function with a scaling parameter
of 0.89±0.03 and an offset of 0.030+0.004
−0.003 . We confirm the
lack of the lowest-mass planets orbiting metal-rich stars and
we also find a period-metallicity correlation for giant planets.
The population of planets with masses >0.1 MJ and orbital
periods less than 100 days is found to be more metal-rich
than the same mass planets with orbital periods greater than
100 days. The difference is significant at the 0.004% level
and the mean difference is found to be 0.16 dex between the
two populations. This result could be describing the forma-
tion locations of planets in the early disks, with metal-rich
disks forming planets in the inner regions and metal-poor
disks forming planets further out in the disk, or that giant
planets migrate more in metal-rich disks, due to a stronger
torque interaction between the high surface density disk and
the migrating planet.
ACKNOWLEDGMENTS
We thank the anonymous referee for providing a fair and
helpful report. We are also thankful for the useful dis-
cussions with Francois Menard and to Andres Jord´an for
providing access to the CORALIE pipeline. JSJ acknowl-
edges funding by Fondecyt through grants 1161218 and
3110004, partial support from CATA-Basal (PB06, Coni-
cyt), the GEMINI-CONICYT FUND and from the Comit´e
Mixto ESO-GOBIERNO DE CHILE, and from the Coni-
cyt PIA Anillo ACT1120. MJ acknowledges financial sup-
18
J.S. Jenkins et al.
port from Fondecyt project #3140607. DM is supported by
the BASAL CATA Center for Astrophysics and Associated
Technologies through grant PFB-06, by the Ministry for the
Economy, Development, and Tourisms Programa Iniciativa
Cientfica Milenio through grant IC120009, awarded to Mille-
nium Institute of Astrophysics (MAS), and by FONDECYT
Regular grant No. 1130196. This research has made use of
the SIMBAD database and the VizieR catalogue access tool,
operated at CDS, Strasbourg, France.
REFERENCES
Anglada-Escud´e G., L´opez-Morales M., Chambers J. E., 2010,
ApJ, 709, 168
Anglada-Escud´e G., et al., 2013, A&A, 556, A126
Arriagada P., Butler R. P., Minniti D., L´opez-Morales M., Shect-
man S. A., Adams F. C., Boss A. P., Chambers J. E., 2010,
ApJ, 711, 1229
Bakos G. ´A., Csubry Z., Penev K., et al. 2012, preprint,
(arXiv:1206.1391)
Balmforth N. J., Korycansky D. G., 2001, MNRAS, 326, 833
Baranne A., et al., 1996, A&A, 119, 373
Brahm R., Jordan A., 2016, in prep
Buchhave L. A., Latham D. W.,
2015,
preprint,
(arXiv:1507.03557)
Butler R. P., Marcy G. W., Williams E., McCarthy C., Dosanjh
P., Vogt S. S., 1996, PASP, 108, 500
Butler R. P., Marcy G. W., Vogt S. S., Fischer D. A., Henry
G. W., Laughlin G., Wright J. T., 2003, ApJ, 582, 455
Butler R. P., et al., 2006, ApJ, 646, 505
Dawson R. I., Fabrycky D. C., 2010, ApJ, 722, 937
Demarque P., Woo J.-H., Kim Y.-C., Yi S. K., 2004, ApJS,
155, 667
Dumusque X., et al., 2012, Nature, 491, 207
Ercolano B., Clarke C. J., 2010, MNRAS, 402, 2735
Fischer D. A., Valenti J., 2005, ApJ, 622, 1102
Foreman-Mackey D., Hogg D. W., Lang D., Goodman J., 2013,
PASP, 125, 306
Haario H., Saksman E., Tamminen J., 2001, Bernoulli, 7, 223
Haario H., Laine M., Mira A., Saksman E., 2006, Statistics and
Computing, 16, 339
Hatzes A. P., 2002, Astronomische Nachrichten, 323, 392
Hatzes A. P., 2013, ApJ, 770, 133
Høg E., et al., 2000, A&A, 355, L27
Ida S., Lin D. N. C., 2004, ApJ, 604, 388
Isaacson H., Fischer D., 2010, ApJ, 725, 875
Ivanyuk O., Jenkins J., Pavlenko Y., Jones H., Pinfield D., 2016,
MNRAS, submitted
Jenkins J. S., Tuomi M., 2014, ApJ, 794, 110
Jenkins J. S., et al., 2006, MNRAS, 372, 163
Jenkins J. S., Jones H. R. A., Pavlenko Y., Pinfield D. J., Barnes
J. R., Lyubchik Y., 2008, A&A, 485, 571
Jenkins J. S., et al., 2009, MNRAS, 398, 911
Jenkins J. S., et al., 2011a, in European Physical Journal Web of
Conferences. p. 2004, doi:10.1051/epjconf/20111602004
Jenkins J. S., et al., 2011b, A&A, 531, A8
Jenkins J. S., et al., 2013a,
in European Physical Jour-
nal Web of Conferences. p. 5001 (arXiv:0912.5325),
doi:10.1051/epjconf/20134705001
Jenkins J. S., et al., 2013b, ApJ, 766, 67
Jenkins J. S., Tuomi M., Brasser R., Ivanyuk O., Murgas F.,
Jones H. R. A., Butler R. P., Tinney C. G., Marcy G. W., Carter
B. D., Penny A. J., McCarthy C., Bailey J., 2006, MNRAS,
369, 249
Jones M. I., Jenkins J. S., Bluhm P., Rojo P., Melo C. H. F.,
2014, A&A, 566, A113
Jones M. I., et al., 2016, A&A, 590, A38
Jord´an A., Brahm R., Bakos G. ´A., et al. 2014, AJ, 148, 29
Kretke K. A., Lin D. N. C., 2012, ApJ, 755, 74
Laughlin G., Steinacker A., Adams F. C., 2004, ApJ, 608, 489
Li H., Lubow S. H., Li S., Lin D. N. C., 2009, ApJL, 690, L52
Lo Curto G., et al., 2010, A&A, 512, A48
Lopez S., Jenkins J. S., 2012, ApJ, 756, 177
L´opez-Morales M., et al., 2008, AJ, 136, 1901
Marcy G. W., Butler R. P., 1996, in Proc. SPIE Vol. 2704, p.
46-49, The Search for Extraterrestrial Intelligence (SETI) in
the Optical Spectrum II, Stuart A. Kingsley; Guillermo A.
Lemarchand; Eds.. pp 46–49
Marcy G. W., Butler R. P., Fischer D., Vogt S. S., Lissauer J. J.,
Rivera E. J., 2001, ApJ, 556, 296
Marsakov V. A., Shevelev Y. G., 1988, Bulletin d'Information du
Centre de Donnees Stellaires, 35, 129
Marsh T. R., 1989, PASP, 101, 1032
Marshall J. P., Moro-Mart´ın A., Eiroa C., et al. 2014, preprint,
(arXiv:1403.6186)
Mayor M., Queloz D., 1995, Nature, 378, 355
Meschiari S., Wolf A. S., Rivera E., Laughlin G., Vogt S., Butler
P., 2009, PASP, 121, 1016
Minniti D., Butler R. P., L´opez-Morales M., Shectman S. A.,
Adams F. C., Arriagada P., Boss A. P., Chambers J. E., 2009,
ApJ, 693, 1424
Mizuno H., 1980, Progress of Theoretical Physics, 64, 544
Mordasini C., Alibert Y., Benz W., Klahr H., Henning T., 2012,
preprint, (arXiv:1201.1036)
Mortier A., Santos N. C., Sozzetti A., Mayor M., Latham D.,
Bonfils X., Udry S., 2012, A&A, 543, A45
Murgas F., Jenkins J. S., Rojo P., Jones H. R. A., Pinfield D. J.,
2013, A&A, 552, A27
Nelson R. P., 2005, A&A, 443, 1067
Nelson R. P., Papaloizou J. C. B., 2004, MNRAS, 350, 849
Nelson B. E., Robertson P. M., Payne M. J., Pritchard S. M.,
Deck K. M., Ford E. B., Wright J. T., Isaacson H. T., 2016,
MNRAS, 455, 2484
Newton M. A., Raftery A. E., 1994, JRSS.B, 56, 3
O'Toole S. J., Tinney C. G., Jones H. R. A., Butler R. P., Marcy
G. W., Carter B., Bailey J., 2009, MNRAS, 392, 641
Pavlenko Y. V., Jenkins J. S., Jones H. R. A., Ivanyuk O., Pinfield
D. J., 2012, MNRAS, p. 2643
Pepe F., Mayor M., Galland F., Naef D., Queloz D., Santos N. C.,
Udry S., Burnet M., 2002, A&A, 388, 632
Perryman M. A. C., Lindegren L., Kovalevsky J., et al. 1997,
A&A, 323, L49
Pinotti R., Arany-Prado L., Lyra W., Porto de Mello G. F., 2005,
MNRAS, 364, 29
Pojmanski G., 1997, Acta Astronomica, 47, 467
Queloz D., et al., 2001, A&A, 379, 279
Rajpaul V., Aigrain S., Osborne M. A., Reece S., Roberts S. J.,
2015, preprint, (arXiv:1506.07304)
Reffert S., Bergmann C., Quirrenbach A., Trifonov T., Kunstler
A., 2015, A&A, 574, A116
Robertson P., Mahadevan S., Endl M., Roy A., 2014, Science,
345, 440
Saar S. H., Donahue R. A., 1997, ApJ, 485, 319
Santos N. C., Israelian G., Mayor M., 2004, A&A, 415, 1153
Santos N. C., Gomes da Silva J., Lovis C., Melo C., 2010, A&A,
2013c, ApJ, 771, 41
511, A54
Jenkins J. S., Yoma N. B., Rojo P., Mahu R., Wuth J., 2014,
MNRAS, 441, 2253
Santos N. C., et al., 2013, A&A, 556, A150
Soto M. G., Jenkins J. S., Jones M. I., 2015, MNRAS, 451, 3131
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
19
Figure A1. HARPS radial velocity timeseries of HD72673 with
the mean subtracted off the data. The upper panel shows the full
timeseries and the lower panel is a zoom in on the most densely
observed epoch for this star.
Sousa S. G., Santos N. C., Israelian G., Mayor M., Udry S., 2011,
A&A, 533, A141
Sozzetti A., 2004, MNRAS, 354, 1194
Takeda G., Ford E. B., Sills A., Rasio F. A., Fischer D. A., Valenti
J. A., 2007, ApJS, 168, 297
Tamuz O., et al., 2008, A&A, 480, L33
Tuomi M., 2012, A&A, 543, A52
Tuomi M., 2014, MNRAS, 440, L1
Tuomi M., Anglada-Escud´e G., 2013, A&A, 556, A111
Tuomi M., et al., 2013, A&A, 551, A79
Tuomi M., Jones H. R. A., Barnes J. R., Anglada-Escud´e G.,
Jenkins J. S., 2014, MNRAS, 441, 1545
Udry S., et al., 2007, A&A, 469, L43
Valenti J. A., Fischer D. A., 2005, ApJS, 159, 141
Wittenmyer R. A., et al., 2012, ApJ, 753, 169
Yasui C., Kobayashi N., Tokunaga A. T., Saito M., Tokoku C.,
2009, ApJ, 705, 54
Figure A2. Time-series and mean subtracted radial velocities for
the star HD72673. The dashed line represents the best straight
line fit to the data. The rms scatter around the fit in m s−1is
shown in the plot.
ual radial velocity measurements where we have removed
5σ outliers that corresponded to bad weather observations,
and therefore had very low S/N data, like the point at BJD-
2453724.79378. After subtracting off the mean of the data,
which we use as our standard flat noise model, we find a
rms of 1.44 m s−1. The lower panel shows the same data
except zoomed in on the most densely sampled observing
epoch (BJD 2453500 - 2454000). Some structure is found in
the radial velocities throughout this epoch and could be the
first signatures of low-amplitude Doppler shifts induced by
orbiting low-mass planets, or stellar activity signals affect-
ing the velocities. In any case the HARPS velocities agree
that HD72673 does not show large radial velocity variations
and is a useful star for testing the precision we can achieve
with our CORALIE pipeline.
Yu C., Li H., Li S., Lubow S. H., Lin D. N. C., 2010, ApJ, 712, 198
van Leeuwen F., 2007, A&A, 474, 653
A1 CORALIE Observations
APPENDIX A: STABLE STAR RESULTS
To confirm the stability of the CORALIE pipeline reduction
and analysis procedure we have observed a star known to
be radial velocity stable at the ∼2 m s−1level that should
provide an ideal test candidate for our method. HD72673 is a
bright (V =6.38), nearby (12.2 pc), and inactive (logR′
HK = -
4.946 Isaacson & Fischer 2010) G9 dwarf star that has a
metallicity of -0.38±0.04 dex (Marsakov & Shevelev 1988;
Santos et al. 2004; Valenti & Fischer 2005), and a mass and
age of 0.814±0.032 M⊙ and 1.48+5.44
−1.48 Gyrs (Takeda et al.
2007), respectively.
In the upper panel of Fig. A1 we show the radial ve-
locity timeseries for HD72673 observed with HARPS that
was taken from the ESO Archive1. The data span a base-
line of over 1900 days in total and comprise 363 individ-
1 Based on data obtained from the ESO Science Archive Facility
under request number JJENKINS 50958.
MNRAS 000, 1–20 (2002)
Over the course of four years we have performed 108 observa-
tions of the star HD72673 with CORALIE, combining data
from this project and also from the HAT-South (Bakos et al.
2012) CORALIE observations. The sampling and time base-
line provide an excellent diagnostic test of the long term sta-
bility that is currently attained with the CORALIE using
the procedure described in Jord´an et al. (2014).
For the observations of HD72673 we aimed to get a S/N
of around 100 across the optical regime of interest, leading
to typical integration times of ∼5 minutes. Fig. A2 shows
the full radial velocity dataset as a function of time and
clearly we see only a small linear trend over the full base-
line of observations. No large systematic trends are found
in our dataset and the gradient of the best fit we show
is 0.004 m s−1/day, well below the intrinsic scatter of our
procedure. The rms scatter for the full data is found to be
10.9 m s−1, however after removing a 5-σ outlier due to low
S/N we arrive at a scatter of 8.7 m s−1, or 8.6 m s−1after
subtraction of the linear trend shown in the figure. There-
fore, we consider the precision of the CORALIE observa-
tions to be 9 m s−1, consistent with the precision reported
by Jord´an et al., but covering a longer time baseline.
20
J.S. Jenkins et al.
APPENDIX B: MMSE PERIODOGRAMS
The minimum mean square error periodograms from the ra-
dial velocity timeseries data discussed in this work. The ver-
tical dashed lines mark the signals detected by the MCMC
search algorithm.
APPENDIX E: ACTIVITY INDICATOR
PERIODOGRAMS
Lomb-Scargle periodograms of the activity indicators mea-
sured from the CORALIE and HARPS timeseries spectra.
APPENDIX F: ASAS PERIODOGRAMS
Lomb-Scargle periodograms of the ASAS timeseries V -band
photometric data for the six stars that show peaks that could
be related to magnetic activity on the surface of the star.
APPENDIX G: RADIAL VELOCITIES
Here we provide all radial velocities that are discussed in
this work.
APPENDIX C: MCMC POSTERIOR
DENSITIES
Here we show the posterior densities from our MCMC search
for signals in the radial velocities for all targets in this work.
We show the densities from the samplings for the periods,
semiamplitudes, and eccentricities for all signals.
APPENDIX D: CORALIE CHROMOSPHERIC
ACTIVITY INDICES
We measure the CORALIE activities using only four echelle
orders, even though the regions we require for the SMW pass-
bands are found across five orders. We drop one of the orders
(order 4) due to an excess of noise at the blue end, which
is due to the position of the echellogram where the V pass-
band is found and therefore including this order enhances
the uncertainty in the S-index and, in general, artificially
increases the activity value making each star appear more
active than it really is. We note that this could be taken out
by calibration to other chromospheric indexes. We show the
CORALIE extraction regions in Fig. D1.
We compute the activities by integrating the square
continuum V (3891-3911A) and R (3991-4011A) bandpasses
and taking the ratio of these against the integrated flux in
the triangular core bandpasses, described in the following
series of equations:
fj,i = ℑj,i ∗ Bj,i ∗ δλ
(D1)
Scont = P (fV,i + fR,i)
P (BV,i + BR,i) ∗ δλ
Score = P (fK,i + fH,i)
P (BK,i + BH,i) ∗ δλ
rP(cid:16)q(σ2
)(cid:17)2
P (BV,i + BR,i) ∗ δλ
rP(cid:16)q(σ2
)(cid:17)2
P (BK,i + BH,i) ∗ δλ
+ σ2
fK,i
+ σ2
fH,i
fV,i
fR,i
(D2)
(D3)
(D4)
(D5)
σScont =
σScore =
The cont and core subscripts represent the continuum
and core regions of the spectrum respectively, ℑ is the
flux measured in each wavelength domain (i), j here de-
notes either the V,R,K, or H bandpass regions, and δλ is
the wavelength step (dispersion), which at the resolution of
CORALIE is ∼0.023A.
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
21
Figure B1. Periodograms from top left to bottom right: HD9174, HD48265, HD68402, HD72892, HD128356, HD143361, HD147873,
HD147873 residuals after fitting out the best fit Keplerian signal associated with the raw data primary spike, HD152079, HD154672,
HD165155, and HD224538. The dashed vertical lines represent the periods detected by the MCMC analysis.
Table G1. HD9174
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
2455160.5362269
2455161.5372730
2455162.5377306
2455467.6819298
2455468.7578863
2455878.6618959
2455879.6756525
2456160.7988170
2456164.7903294
2456307.5567717
2456308.5698681
2456675.6083797
2456676.6025593
2456881.8675908
2456882.8265879
2456883.7721760
7.2
-11.8
30.2
-35.8
-25.8
-8.9
-5.9
6.0
38.0
23.0
20.0
-22.0
-16.0
-20.2
-17.2
-13.2
9.0
9.0
9.0
9.0
9.0
9.0
15.0
9.0
9.0
9.0
9.0
9.0
9.0
12.0
12.0
12.0
2455188.6187049
2455883.6145165
2455885.5904695
2456183.7872355
2456184.6697746
2456184.8402196
2456185.7970768
2456442.9352036
2456461.9331703
2456463.8950719
2456561.7814364
2456562.5748725
2456563.6078357
10.77
-10.66
-12.54
12.43
9.39
8.10
13.90
3.98
0.26
1.48
-13.46
-12.24
-13.76
0.55
0.86
1.03
0.48
0.79
0.92
0.55
0.69
0.74
0.77
0.50
0.86
0.78
MNRAS 000, 1–20 (2002)
22
J.S. Jenkins et al.
Figure C1. Posterior densities for the signals HD9174b, HD48265b, HD68402b, HD72892b, and HD128356b.
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
23
Figure C2. Posterior densities for the signals HD143361b, HD147873b, HD147873c, HD152079b, and HD154672b.
MNRAS 000, 1–20 (2002)
24
J.S. Jenkins et al.
Figure C3. Posterior densities for the signals HD165155b and HD224538b.
Figure D1. CORALIE echelle orders used to extract the S activity indices for HD128356. The square continuum bandpasses regions
V and R are shown by the thick solid lines in the top and bottom plots, respectively. The thick triangular solid lines in the upper and
lower center plots show the K and H bandpasses, respectively.
MNRAS 000, 1–20 (2002)
New Planetary Systems from the CHEPS
25
Figure E1. In each plot from top to bottom we show the periodograms for the BIS, FWHM, S, Hα, and He I indices. From top left
to bottom right we show the stars HD9174, HD48265, HD68402, HD72892, HD128356, HD143361, HD147873, HD152079, HD154672,
HD165155, and HD224538, respectively.
MNRAS 000, 1–20 (2002)
26
J.S. Jenkins et al.
Figure F1. From top left to bottom right we show the periodograms from ASAS data for the stars HD68402, HD147873, HD152079,
HD154672, HD165155, and HD224538, respectively.
Table G2. HD48265
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
BJD
MIKE
RV
Error
2455208.6992489
2455209.7264368
2455210.7165285
2455465.8329696
2455468.8765488
2455877.7963300
2455878.7843441
2455879.7984724
2455969.6052648
2455970.7499814
2455971.7266404
2456034.5026492
2456037.5096202
2456164.8925279
2456378.6064455
2456379.5444957
2456380.6001445
2456381.5979512
2456733.6079032
2456736.5950580
2456753.5245324
2456754.5405659
1.8
10.8
-29.2
16.8
27.8
-4.2
13.8
10.8
-19.2
-17.2
-21.2
13.8
37.8
15.8
-15.2
-21.2
-6.2
-15.2
2.9
19.9
3.9
7.9
9.0
9.0
9.0
9.0
9.0
15.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
10.0
10.0
10.0
2454365.8248037
2454366.8740710
2454367.8521212
2454580.5231386
2454581.5537085
2454724.8337096
2454725.8177884
2454726.7789038
2455651.5366350
2455883.7630397
2455885.7426863
2455992.5816638
2455994.5576882
2456183.8678319
2456185.8516258
2456442.4578668
2456444.4478133
2456562.8313800
-32.25
-25.18
-29.87
22.43
23.60
-14.14
-13.78
-14.84
-15.41
-7.88
-1.10
18.09
23.81
52.75
57.16
-16.22
-14.86
-12.48
0.77
0.99
0.62
0.64
0.56
0.59
0.53
0.80
0.56
0.70
0.92
0.56
0.61
0.59
0.66
0.91
1.07
0.73
2452920.8629
2453431.6073
2453455.5573
2453685.8138
2453774.6744
2453775.6763
2453784.6887
2453811.5943
2453987.9168
2454078.7826
2454081.7133
2454137.6436
2454138.6670
2454189.5669
2454483.6148
2454501.6725
2454522.6219
20.6
-30.2
-35.0
15.0
25.0
12.4
25.4
24.3
-7.7
-17.9
-23.4
-20.7
-28.7
-29.5
28.5
21.5
25.4
6.9
2.5
2.6
2.9
3.3
5.2
2.6
2.5
2.9
3.0
2.8
2.5
2.6
4.4
3.0
2.4
2.7
Table G3. HD68402
HARPS
RV
-21.04
-17.13
12.99
13.44
11.72
Error
1.51
1.66
1.46
1.51
2.07
BJD
RV
Error
BJD
CORALIE
2456442.5125503
2456444.4892384
2456562.9026514
2456563.8260976
2456563.9098232
2455268.5536332
2455269.5471815
2455270.5544347
2456034.5316938
2456037.5405384
2456307.7270563
2456308.7455901
2456675.7326419
2456675.7405950
2456735.5968024
2456736.6337750
2456752.5562350
2456754.5693098
2456823.4935737
2456825.4656166
2457075.6564372
2457318.8514672
11.4
0.4
4.4
-45.6
-31.6
-32.6
-18.6
47.4
64.4
45.0
33.0
38.0
53.0
33.0
8.0
-43.1
-19.1
9.0
9.0
9.0
9.0
9.0
9.0
9.0
16.0
17.0
12.0
14.0
13.0
15.0
15.0
12.0
14.0
14.0
MNRAS 000, 1–20 (2002)
Table G4. HD72892
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
New Planetary Systems from the CHEPS
27
2456442.5223438
2456443.5281580
2456444.5101783
2456448.4605785
2456449.4624615
2456450.4614518
2456462.4520834
2456463.4499736
-195.55
-167.05
-134.83
144.30
248.19
357.16
-114.51
-137.72
1.12
1.77
1.36
1.29
0.82
1.21
1.13
1.69
2455698.5075694
2455699.5368477
2455700.5250521
2455968.7197064
2455969.7048152
2455970.6308925
2455972.6288495
2456034.5646797
2456037.5717726
2456307.7713660
2456308.7833076
2456378.6598031
2456379.5959537
2456463.4576754
2456464.4530818
2456465.4567122
2456467.4802779
2456675.7562736
2456734.6035167
2456752.6156272
2456754.5855485
2456824.4911691
2457184.4724902
223.8
337.8
444.8
-111.2
-77.2
-55.2
61.8
-129.2
-156.2
-110.3
-126.3
159.7
128.7
-57.3
-80.3
-111.3
-147.3
-196.3
134.1
-176.9
-154.9
-138.9
-161.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
14.0
9.0
16.0
9.0
11.0
13.0
13.0
15.0
13.0
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
2455268.9086205
2455269.8196218
2455349.6784180
2455350.6814856
2455351.6612024
2455352.6667444
2455609.8365887
2455611.8334577
2455699.7015878
2455967.8551252
2455968.8240257
2455970.8305273
2455971.7742759
2455972.8507044
2456034.8259987
2456037.8192588
2456160.4758000
2456164.4757457
2456381.7122851
2456463.5427541
2456464.6240954
2456465.6419917
2456467.5931893
2456467.6621264
2456676.8044908
2456734.7225188
2456752.8173417
2456823.7074373
2456825.7157702
2456881.5522933
2456882.5702439
-22.1
-34.1
-16.1
14.9
-4.1
-6.1
-27.1
-13.1
9.9
27.9
39.9
20.9
31.9
46.9
11.9
27.9
-6.1
-9.1
15.9
-23.1
-35.1
-45.1
-26.1
-24.1
19.8
-35.2
-30.2
1.8
6.8
4.8
20.8
9.0
9.0
9.0
9.0
9.0
9.0
14.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
16.0
9.0
9.0
10.0
12.0
14.0
12.0
12.0
15.0
2454248.6023117
2454248.6108994
2454248.6176931
2454367.5031968
2454369.5091063
2454577.6158842
2454578.5680249
2454581.7357981
2455271.8690903
2455649.7321937
2455650.7148680
2455651.7295683
2455786.4827996
2455787.4566297
2455992.7730262
2455993.7048066
2456063.6469346
2456064.6145915
2456065.6564584
2456183.4873901
2456442.6772101
2456443.6828424
2456444.6629996
2456450.4938930
2456462.6290393
2456463.5862333
2456562.4755850
16.75
21.79
17.99
-31.24
-37.97
35.96
35.50
41.67
-23.11
-6.46
-6.51
-7.26
50.87
49.97
9.21
9.29
34.67
35.94
36.26
-31.26
-46.63
-46.21
-45.18
-40.30
-36.01
-35.50
-3.37
1.66
1.69
0.69
0.66
2.49
0.48
0.52
0.47
0.36
0.46
0.49
0.50
0.51
0.36
0.39
0.52
0.78
0.45
0.43
0.60
0.46
0.77
0.91
0.62
0.54
0.80
0.56
Table G5. HD128356
MNRAS 000, 1–20 (2002)
28
J.S. Jenkins et al.
Table G6. HD143361
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
BJD
2454253.7711993
2454253.7786414
2454367.5565354
2454368.5376348
2454578.7885789
2454581.8076461
2455271.8809609
2455649.7540214
2455786.5132754
2455992.7979663
2455993.7269629
2456063.6718947
2456064.6371793
2456065.6805773
2456184.5078021
2456442.6901584
2456443.6959898
2456444.7117386
2456462.6564547
10.50
9.06
-23.81
-23.64
-83.52
-81.90
19.62
-83.51
-30.99
55.69
65.78
68.27
65.24
64.11
53.36
-22.67
-19.26
-20.46
-25.82
3.17
2.22
1.09
1.13
0.83
0.92
0.77
0.94
0.89
1.23
0.89
2.14
0.87
0.80
0.84
0.90
1.96
1.79
0.97
2452864.57934
2453130.83712
2453872.73696
2453987.50506
2453988.49460
2454190.80550
2454217.84734
2454277.65819
2454299.55951
2454300.58038
2454339.50049
2454501.87197
2454650.68383
2454925.87115
2454963.75707
2454965.78744
2455019.67861
2455269.8328453
2455269.8477334
2455349.6897288
2455349.7090115
2455350.6957245
2455350.7085995
2455351.6772114
2455351.6921046
2455352.6804150
2455352.6925204
2455465.4819439
2455465.4921698
2455466.4810091
2455467.4908225
2455786.5181546
2455787.5128697
2455788.5005189
2455967.8637123
2455969.8564628
2455970.8440395
2455971.8236578
2455972.8379866
2456034.8111668
2456037.8046599
2456160.4914594
2456161.4914636
2456162.5761051
2456164.4909267
2456381.7216801
2456381.7325023
2456463.6056985
2456465.6944440
2456467.5744277
2456554.5120400
2456554.5219817
2456555.5573278
2456676.8679860
2456734.7382463
2456752.8000194
2456754.7548507
2456823.7338512
2456825.7551623
2456881.6125148
2456882.6036411
2457281.5366030
31.9
30.9
9.9
6.9
-5.2
-11.2
3.9
-0.2
-8.2
2.9
-37.2
-41.2
-44.2
-36.2
-28.2
-32.2
-18.2
48.8
55.8
49.8
61.8
52.8
73.8
73.7
49.7
58.7
53.7
46.7
-14.3
-13.3
-23.3
-22.3
-75.3
-28.3
-49.3
-33.3
-68.4
-61.4
-39.4
-51.4
-33.4
-36.4
5.6
17.6
71.5
12.0
12.0
14.0
15.0
17.0
22.0
13.0
14.0
13.0
12.0
19.0
18.0
20.0
13.0
13.0
12.0
12.0
14.0
14.0
14.0
14.0
14.0
14.0
12.0
11.0
12.0
18.0
12.0
15.0
14.0
12.0
12.0
23.0
16.0
20.0
13.0
14.0
12.0
13.0
13.0
13.0
13.0
12.0
15.0
19.0
MIKE
RV
43.03
-6.28
28.04
38.49
25.06
0.00
-7.02
-29.11
-37.73
-36.03
-52.16
-94.15
-105.06
21.04
34.93
31.39
27.65
Error
7.14
2.65
2.62
3.45
2.81
3.43
3.38
3.43
3.04
2.51
3.39
3.01
6.82
2.76
3.05
2.82
2.46
MIKE
RV
0.00
-9.92
-5.91
-72.79
134.00
156.04
Error
3.38
3.36
3.56
5.00
4.93
4.48
HARPS
RV
-165.09
86.26
77.96
-117.74
-121.37
-120.54
124.30
-73.82
-102.21
-93.49
-92.88
-86.93
-91.50
-83.90
-77.92
-80.38
-90.35
250.12
258.36
223.49
219.57
95.12
62.78
Error
BJD
2453189.66944
2453190.64093
2453191.65473
2453551.62308
2454339.53949
2455001.70315
4.83
2.48
1.73
1.27
1.56
1.57
1.42
2.01
2.08
1.66
2.53
1.92
2.10
3.01
1.67
2.00
2.11
3.48
1.87
3.03
3.41
2.19
3.16
Table G7. HD147873
CORALIE
BJD
2455268.8296124
2455269.8954986
2455270.8623072
2455349.7478270
2455351.7263025
2455352.7243194
2455465.5203302
2455466.5103888
2455467.5077317
2455468.4926389
2455786.5547221
2455787.5512296
2455788.5404119
2455968.8355111
2455970.8644298
2456034.8536087
2456037.8611090
2456160.5035894
2456161.5099351
2456164.5890537
2456307.8687079
2456308.8690886
2456381.7577061
2456463.6629499
2456464.6620073
2456554.5635006
2456555.4998324
2456676.8378428
2456733.8400326
2456734.7541033
2456735.7646576
2456736.8330743
2456752.8898318
2456754.7704681
2456823.7610291
2456824.7569383
2456825.8288093
2456881.6254682
2456882.6396241
2456883.5912828
2457179.6595698
2457180.6006537
2457181.5978747
2457182.5980141
2457183.7046858
2457184.6231976
RV
66.7
58.7
41.7
4.7
-9.3
3.7
38.7
25.7
39.7
32.7
-195.3
-190.3
-170.3
100.7
74.7
-19.3
8.7
15.7
22.7
55.7
112.7
117.7
-119.3
-189.3
-190.3
144.7
118.7
-0.3
-56.8
-50.8
-55.8
-47.8
50.2
76.2
-155.8
-181.8
-168.8
130.2
151.2
149.2
-126.8
-127.8
-142.8
-120.8
-119.8
-111.8
Error
BJD
2454365.5676419
2454580.8338639
2454581.8459239
2454724.4802846
2454725.4983255
2454726.4832450
2455271.8407105
2455649.7654997
2455786.5492317
2455787.4838577
2455787.6993011
2455788.4587689
2455788.6109888
2455788.7074614
2455992.8338522
2455993.7871396
2455994.7989047
2456063.7089752
2456064.7231089
2456183.4686863
2456184.4750146
2456442.7882798
2456444.7375998
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
12.0
11.0
11.0
12.0
14.0
13.0
14.0
14.0
14.0
14.0
13.0
13.0
16.0
15.0
13.0
13.0
14.0
13.0
MNRAS 000, 1–20 (2002)
Table G8. HD152079
New Planetary Systems from the CHEPS
29
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
BJD
MIKE
RV
Error
2455786.6489688
2456463.6745808
2456464.7049704
2456465.6766024
2456555.5745883
2456676.8549118
2456733.8533258
2456734.8495354
2456735.8130233
2456736.8653294
2456752.8688188
2456754.7851172
2456823.7916933
2456824.7696220
2456825.8154610
2456881.6859346
2456882.6527789
-41.7
14.3
17.3
8.3
-0.8
2.2
-1.2
-17.2
-10.2
7.8
8.8
0.8
16.8
-3.2
-8.2
10.8
-10.2
9.0
9.0
9.0
9.0
9.0
15.0
13.0
11.0
12.0
15.0
12.0
14.0
13.0
13.0
13.0
13.0
12.0
2454253.8065363
2454367.5811986
2454579.8264105
2455649.8008277
2455650.7595904
2455651.8003619
2455786.5261504
2455787.5966780
2455992.9178275
2455993.8444865
2456064.7106335
2456184.5322797
2456442.7259868
2456443.7319729
2456444.7714764
2456462.7752093
2456561.5772596
2456563.5044073
-20.75
-22.95
-29.21
-30.70
-29.43
-29.35
-22.51
-22.46
1.25
3.09
19.86
44.23
24.77
27.81
25.93
19.99
18.97
18.41
1.75
1.00
0.72
1.01
0.96
0.84
0.96
0.85
1.01
0.93
0.89
0.93
0.90
1.62
1.28
1.30
1.25
1.18
2452917.4972
2453542.6649
2453872.8022
2453987.5436
2453988.5202
2454190.8274
2454277.6950
2454299.6134
2454725.5353
2454925.9161
2454963.7753
2454993.7093
2455001.7291
2455017.6624
2455019.6938
-24.3
22.5
-8.5
-10.3
-12.6
-13.7
-19.7
-19.6
-35.1
-29.2
-22.6
-27.5
-25.3
-28.5
-22.5
6.2
3.3
2.5
2.8
2.7
2.9
3.4
3.3
2.6
2.4
2.7
2.4
2.9
2.4
2.2
HARPS
RV
225.90
227.85
82.73
71.32
-18.52
-195.00
-87.04
-126.00
-98.02
53.14
50.37
48.70
188.65
-86.65
-86.90
-163.39
-164.91
39.66
36.30
Error
BJD
0.69
1.10
0.61
0.52
0.63
0.66
0.66
0.78
0.83
1.25
0.66
0.74
0.85
0.69
1.29
0.91
1.02
0.85
0.89
2453189.71323
2453190.70833
2453191.72040
2453254.50616
2453596.68926
2453810.90968
2453872.81362
2454189.87086
2454189.87898
2454190.84022
2454215.86050
2454216.78927
2454217.87254
2454277.70250
2454299.62096
2454339.55738
2454501.89596
2455018.68932
MIKE
RV
-141.83
-143.62
-147.14
175.30
130.45
-5.00
260.75
-80.09
-90.28
-57.35
242.59
243.77
250.94
67.40
0.00
-145.72
-143.22
213.92
Error
2.89
2.73
3.33
2.60
3.10
2.56
2.45
3.62
3.52
2.84
2.52
2.54
2.76
2.74
2.64
4.15
2.71
2.17
Table G9. HD154672
BJD
RV
Error
BJD
CORALIE
2454367.5951401
2454368.6185060
2454578.8112608
2454581.8703510
2455271.8920649
2455649.8248173
2455786.5972307
2455993.8674482
2455994.8361631
2456063.7424430
2456064.7465191
2456065.7448182
2456183.5270912
2456442.7630531
2456443.7800510
2456462.7983801
2456462.8102434
2456561.4986867
2456563.5280947
2455268.8925184
2455286.7126041
2455286.7482435
2455287.7447603
2455287.7803973
2455288.7337184
2455288.7776633
2455349.7906916
2455350.8687367
2455351.7067623
2455352.7571340
2455353.6007870
2455354.6329905
2455354.7244421
2455355.6194464
2455355.7115550
2455433.5945428
2455434.5939320
2455467.5419737
2455468.5360292
2455786.6672674
2456034.8785575
2456037.8864541
2456160.5115180
2456161.5235072
2456164.5556060
2456381.7967325
2456463.6867223
2456464.7194041
2456734.8670344
2456735.8301098
2456823.8058978
2456825.8537984
2456882.6789128
-51.4
-133.4
-137.4
-130.4
-136.4
-139.4
-138.4
161.6
177.6
173.6
172.6
135.6
151.6
136.6
135.6
148.6
-58.4
-68.4
-152.4
-140.4
-128.5
94.5
86.5
-99.5
-68.5
60.5
47.5
-212.5
-226.5
-17.5
-20.5
153.5
178.5
44.5
9.0
12.0
12.0
12.0
12.0
11.0
12.0
10.0
10.0
10.0
9.0
10.0
10.0
11.0
11.0
11.0
11.0
10.0
11.0
11.0
10.0
9.0
9.0
11.0
12.0
13.0
10.0
10.0
10.0
9.0
9.0
11.0
11.0
11.0
Table G10. HD165155
BJD
RV
Error
BJD
CORALIE
HARPS
RV
Error
2455698.7379179
2455786.6994564
2455786.7108277
2455787.6641024
2455787.6754805
2455788.6765746
2455788.6875372
2456034.9332917
2456381.8066463
2456381.8190355
2456463.7613374
2456464.7345595
2456555.5907771
2456555.6041047
2456733.8878823
2456735.8673459
2456754.8071800
2456823.8481613
2456825.8695333
2456881.7083646
2456883.6490320
2457077.8839217
2457183.7828036
2457184.6886555
2457312.5069791
2457318.4944200
-70.4
36.6
75.6
35.6
35.6
48.6
67.6
20.6
54.6
47.6
59.6
11.6
-42.4
-39.4
145.2
111.2
115.2
112.2
89.2
99.2
85.2
120.2
-250.8
-250.8
-254.8
-279.8
9.0
13.0
13.0
12.0
12.0
13.0
13.0
12.0
12.0
12.0
9.0
9.0
11.0
9.0
10.0
10.0
13.0
11.0
11.0
11.0
10.0
13.0
16.0
15.0
17.0
16.0
2454577.8573518
2454578.8259992
2454579.8512477
2455650.8171652
2455786.6216504
2455787.4979049
2455788.6355855
2455993.8895322
2455994.8596255
2456063.7543293
2456064.7595490
2456065.7693045
2456183.5412473
2456442.8006921
2456443.8064819
2456444.8081381
2456461.7756434
2456462.8238912
2456463.7557595
2456561.4867891
70.07
67.18
67.69
-83.60
-5.87
-3.44
-1.89
-6.64
-7.98
-35.54
-35.45
-33.41
-15.93
14.89
19.51
17.67
13.11
9.33
11.89
-62.19
0.70
0.80
0.74
1.14
1.04
0.96
1.07
1.23
1.57
1.82
0.96
2.68
1.13
1.18
2.63
1.57
1.08
1.42
1.71
1.12
MNRAS 000, 1–20 (2002)
30
J.S. Jenkins et al.
Table G11. HD224538
MIKE
RV
-2.66
0.00
7.12
-13.98
3.92
-7.75
Error
2.86
3.27
2.85
3.23
3.86
3.09
BJD
RV
Error
BJD
CORALIE
2455196.5457833
2455197.5258897
2455198.5268449
2455352.9431841
2455465.6728602
2455468.6569334
2455497.5904524
2455514.6207088
2455786.8585090
2455787.8646502
2455788.8218648
2455877.6116650
2455878.5958784
2455879.6063410
2455969.5203878
2455970.5204393
2456160.6541736
2456164.7802416
2456554.6541123
2456554.6624176
2456555.7132051
2456675.5534417
2456824.9242966
2457317.6711775
-63.9
-62.9
-54.9
-89.9
-55.9
-51.9
-52.9
-42.9
37.1
25.1
53.1
93.1
92.1
108.1
158.1
167.1
-25.9
-4.9
-68.9
-54.9
-76.9
-27.9
5.9
-3.1
2454365.6728323
2454724.7162997
2454725.6845580
2454726.6664725
2454727.6026469
2455786.8072676
2455787.6269341
2455787.8105224
2455787.9313791
2455788.5991810
2455788.7826534
2455883.5865245
2455885.5121047
2455885.5744505
2455885.7033086
2456064.9384446
2456183.7248653
2456185.6754003
2456442.9267855
2456461.8887446
2456561.7050322
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
9.0
19.0
22.0
9.0
9.0
16.0
16.0
16.0
16.0
9.0
9.0
9.0
16.0
13.0
36.0
HARPS
RV
-72.92
87.00
87.50
89.44
87.24
-5.84
-12.49
-16.34
-14.21
-14.30
-14.01
46.83
54.40
52.39
51.81
67.21
-67.99
-70.76
-116.37
-110.55
-109.01
Error
BJD
2453189.91295
2453190.91771
2453191.91314
2453254.73446
2454338.83927
2454339.75448
0.84
0.93
1.02
0.65
0.83
0.80
0.96
0.73
0.82
1.26
1.33
1.19
1.30
1.53
1.43
0.85
0.86
1.02
1.14
1.03
0.97
MNRAS 000, 1–20 (2002)
|
1304.6043 | 1 | 1304 | 2013-04-22T18:15:51 | Solid state physics of impact crater formation: further considerations | [
"astro-ph.EP"
] | Impact craters exist on solid surface planets, their satellites and many asteroids.The aim of this paper is to propose a theoretical expression for the product $\rho r^{3} v_{1}^{2}$,where the three symbols denote the mass density,radius and speed of the impactor. The expression is derived using well known results of solid state physics,and it can be used in estimating parameters of impactors which have led to formation of craters on various solid bodies in the Solar System. | astro-ph.EP | astro-ph |
Solid state physics of impact crater formation: further considerations
V.Celebonovic
Institute of Physics,Pregrevica 118,11080 Zemun-Belgrade,Serbia
[email protected]
Abstract
Impact craters exist on solid surface planets, their satellites and
many asteroids.The aim of this paper is to propose a theoretical ex-
2
1,where the three symbols denote the
pression for the product ρr
mass density,radius and speed of the impactor. The expression is de-
rived using well known results of solid state physics,and it can be used
in estimating parameters of impactors which have led to formation of
craters on various solid bodies in the Solar System.1
v
3
1to appear in Rev.Mex.Astron.Astrophys, October 2013
1
Introduction
Craters of various sizes have been observed on the terrestrial planets,their
satellites and the major part of the asteroids.The study of these craters, and
the resulting constraints on the related impactors has become a separate
field of research in planetary science,see for example [8].At the time of this
writing, the latest example testifying about the importance of impacts into
the Earth and their consequences is the small asteroid which entered the
atmosphere over the city of Chelyabinsk in Russia on February 15, 2013.
One of the fundamental questions concerning the impactors is what can be
concluded about them by combining astronomical data with results of solid
state physics.
Recent theoretical work for example [1] has shown that the application
of elementary principles of solid state physics to this problem gives physi-
cally plausible results in reasonable agreement with those obtained by celes-
tial mechanics. It was assumed there that the material of the target was a
crystal lattice,and the calculations were performed per unit volume. It was
assumed in that paper that the condition for the formation of a crater is
that the kinetic energy of a unit volume of the impactor has to be equal
to the internal energy of a unit volume of the material of the target. The
result was an expression for the minimal speed which an impactor of given
parameters must have when hitting a target with a predefined set of pa-
rameters, in order to form a crater. Obviously,a certain part of surfaces of
planets,satellites,asteroids,.. in which impacts occur are granular. For recent
theoretical work on "granular impact" see, for example,[3] or [11].
The aim of the calculation reported here is to take into account the di-
mensional effects - to consider both the impactor and the crater it forms as
objects of finite dimensions. The novelty of the approach discussed in the
next section, compared to existing work such as [4],is the generality - it is
based on principles of solid state physics,and it can be applied to any solid
material. Formation of impact craters is here discussed from the viewpoint
of pure solid state physics as the following analogous problem: what kinetic
energy of an impactor is needed to produce a hole of given dimensions in a
material with a predefined set of parameters? The calculations will be per-
formed under the assumptions that the material of the target is a crystal,that
one of the five usual bonding types exists in it and that in the impact the
target does not heat up to its melting temperature, so that solid state physics
can be applied. For recent work on the problem of heating in impacts see [2].
1
The expression for the cohesion energy used in the present work is one of the
theoretically possible expressions. More details are avaliable, for example, in
[7].
2 Calculations
The physical keypoint of the calculation to be discussed in this section is the
idea that the kinetic energy of the impactor must be greater than or equal
to the internal energy of some volume, V2 of the material of the target. The
kinetic energy of the impactor of mass m1 and speed v1 is obviously:
Ek =
1
2
m1v2
1
(1)
The internal energy of a volume V2 of the target material is equal to the sum
of the following three "contributions" : the cohesion energy EC,the thermal
energy ET and the energy EH (T ) required for heating the specimen by an
amount ∆T in the moment of impact:
EI = EC + ET + EH(T )
(2)
The condition for the formation of a crater as a consequence of impact of a
projectile into a target is, in general terms, given by
Ek = EI
(3)
In order to ensure the stability of matter, the interatomic interaction energy
must have the form
where R is the inetartomic distance, the first term on the right hand side is
attractive and the second one repulsive.It can be shown [5] that the cohesion
energy is given by
where B0 = −V (∂P/∂V )T = ρ(∂P/∂ρ)T is the bulk modulus of a material
and Ωm the volume per pair [5]. The speed of sound waves in a material with
pressure P and density ρ is given by
¯u2 =
∂P
∂ρ
2
(6)
E =
−K
Rp +
C
Rn
EC = −
9B0Ωm
pn
(4)
(5)
and therefore
EC = −9
ΩmB0
pn
= −9(
Ωm
pn
)ρ
∂P
∂ρ
= −9
Ωmρ¯u2
pn
(7)
The analytical form of the function ET depends on the temperature of the
solid. It can be shown (for example [7]) that for temperatures
kBT0 >>
¯h¯u
a
(8)
(¯h is Planck's constant divided by 2π, ¯u is the mean speed of sound in the solid
and a is the lattice constant),one should use the "high temperature" limit
for the thermal energy. Inserting for example ¯u = 5km/s and a = 5 × 10−6m
it follows that T0 >> 77K.
The thermal energy ET is given by
ET = 3NνkBT D3(x)
(9)
where x = TD/T , TD is the Debye temperature,ν is the number of particles
in the elementary crystal cell,N is the number of elementary crystal cells in
the specimen,and kB is Boltzmann's constant. The symbol D3(x) denotes
the n = 3 case of the Debye function, which is given by
Dn(x) =
n
xn Z x
0
tndt
Exp[t] − 1
= (n/xn)xn[
1
n
−
x
2(n + 1)
+
B2kx2k
(2k + n)(2k)!
]
∞
Xk=1
(10)
and B2k denote Bernoulli's numbers. Integrating and retaining the terms up
to and including second order leads to:
ET
∼= 3NνkBT [1 −
3
8
TD
T
+ (
1
20
)(
TD
T
)2]
(11)
The specific heat CV is given by
CV =
∂ET
∂T
= 3NνkB(D3(TD/T ) − (TD/T )D′
3(TD/T ))
(12)
which leads to the following approximate result:
CV
∼= 3NνkB[1 − (
1
20
)(
TD
T
)2 + (
1
560
)(
TD
T
)4]
(13)
3
The energy EH (T ) is obviously given by EH = CV (T1 − T ), where T1 is
the temperature to which the material of the target heats up in the moment
of impact and T is the initial temperature.
One of the problems consists in knowing the volume V2,which is,in first
approximation,equal to the volume of the crater. It is known from theoretical
work [4] and references given there that the form of a crater is determined by
a combination of "gravity scaling" and "strength scaling", where the term
"strength" reffers to the material strength of the target. Explicite expressions
for the form of craters exist in [4] but they are applicable to 4 material types.
In order to simplify somewhat the calculations,and get slightly more general
results, it was assumed in the present work that craters have the shape of
a half of a rotating elipsoid,with distinct semi axes,denoted by a,b and c.
Physically,a and b denote the semi axes of the "opening" of the crater,while
c is the depth. In that approximation,the volume of the crater is obviously
given by
V2 =
πabc
(14)
2
3
Inserting the definition of the mass density (ρ1 = m1
) in eq.(1) and assuming
V1
further that the impactor has the form of a sphere of radius r1,the following
expression for the kinetic energy of the impactor is obtained:
Ek =
1
2
m1v2
1 =
1
2
×
4
3
πr3
1ρ1v2
1 =
2
3
πρ1r3
1v2
1
(15)
Inserting eqs.(7), (11),(13) and (15) into (3) one gets the final form of the
energy condition which has to be satisfied for the formation of an impact
crater:
3kBT1Nν[1 −
+(
1
560
)(
TD
T
)4 −
1
420
)2 +
TD
T1
3
8
T 4
T 3T1
−
−
(
TD
1
20
T
3¯u2ρΩm
npνkBT1
T 2
D
T T1
)
(
1
10
2πρ1
3
] =
r3
1v2
1
(16)
The number N is equal to the ratio of the volume of the crater, approximated
by eq.(14), and the volume of the elementary crystal cell,ve: N = V /ve.
Equation (16) groups known or measurable parameters on the left hand side,
while on the right hand side it contains parameters of the impactor. Applying
this equation demands the knowledge of the equation of state (EOS) of the
material of the target. Generally speaking,regardless of its detailed form, an
4
EOS can be expressed in the following analytical form
P (ρ) =
ai(
ρ
ρ0
)i
∞
Xi=0
(17)
where ai are some coefficients, ρ0 is the density at some pressure P0 and ρ is
the density at pressure P . It follows that
(
∂P
∂ρ
) =
∞
Xi=0
(i + 1)ai+1(
ρ
ρ0
)i
(18)
A well known example of the EOS of a solid is the Birch-Murnaghan EOS
[12]
ρ
ρ0
(19)
3B0
P (ρ) =
ρ
ρ0
)7/3 − (
2 "(
)5/3# ×
n1 + (3/4)(B ′
0 − 4)h(ρ/ρ0)2/3 − 1io
The symbols ρ0 and ρ denote the mass density of a specimen under consid-
eration at the initial value of the pressure P0 and at some arbitrary value
P .
More refined results could be obtained by using analytic approximations
to the Helmholtz free energy, from which all the thermodynamic potentials
can be derived. This is,for example, the approach used in the ANEOS EOS
[13]. However, ANEOS uses a file of up to 40 parameters for the characteri-
zation of materials,which complicates its applications.
3 A test example
As a test of the applicability of the procedure discussed in the preceeding
paragraph, it was applied to the Barringer crater in Arizona. One of the min-
erals found in and around the crater is Forsterite - Mg2SiO4. The dimensions
of the crater are a = b = 1.186km and c = 0.17km. Take that the expo-
nents in the expression for the interparticle potential are p = 1 and n = 9.
Assume furher that ¯V = 5km/s,which is the mean value of the measured
velocity of seismic waves in the Earth, that the temperature before impact is
T = 300K and that the target heats up to T1 = 550K at the point of impact.
This is safely far from the melting temperature of Forsterite (which is around
5
TM = 1900C) so that solid state physics can be applied. All the parameters
of Forsterite exist in the literature, such as [10] or http://webmineral.com.
Taking the necessary values,and assuming that the impactor had a radius of
∼= 41km/s.
r1 = 65m and density of ρ1 = 8500kg/m3 one finally gets that v1
This value is larger than existing results, such as for example [9] or [6].The
discrepancy can be traced to the assumption, made in the calculation re-
ported here, that the target material is only Forsterite. In reality, this is
certainly not the case. Taking that 10 percent of the target is Forsterite,and
∼= 15km/s , which is
that the object had a radius of r1 = 60m would give v1
much closer to the results of celestial mechanics.
4 Discussion
In this paper we have presented a simple procedure which gives the possi-
bility of estimating the value of the product of the density and radius of the
impactor and the speed of impact: ρ1r3
1v2 in terms of various material pa-
rameters of the target and the impactor. The expression for this product has
been derived using basic principles of solid state physics, without any spe-
cial assumption(s) about the materials. In (at least some of the) terrestrial
applications the density of the impactor and the inclination of the trajectory
can (in principle) be measured or estimated, which gives the possibility of
estimating the impact speed if the value of r1 can somehow be estimated.
The shape of the craters has been approximated as a half of a rotating
elipsoid,regardless of the composition of the material of the target. As a
first approximation, the speed of the seismic waves has been "put in" in-
stead of being calculated from the equation of state of the material of the
target. Calculating this speed would have demanded the precise knowledge
0 for
of the chemical composition of the target materials,and of B0 and B ′
them. The impact of a projectile into a target leads to heating,and possi-
bly melting and even vaporisation of the target around the impact point.
The thermodynamic result of an impact depends on the heat capacity of
the material of the target and on the kinetic energy of the impactor. The
heat capacity can be measured, or theoretically estimated, assuming prior
knowledge of the chemical composition of the target. Knowledge of the heat
capacity is vital for estimates of the temperature to which the target heats in
the impact.Therefore, the question is which impacts lead to melting and va-
porisation of the material of the target,and which just provoke heating of the
6
target. Even if the material of the target gets partially melted and as such
flows away, the dimensions of the crater formed by the impact are determined
by the volume of the material pushed away in the impact. This volume is,in
turn, determined by the ratio of the kinetic energy of the impactor to the
internal energy of some volume of the target. Some details of this problem
have been discussed in [2].
The general conclusion of work reported here is that by using simple
well known results of solid state physics and a given appoximation of the
shape of impact craters, it becomes possible to estimate the value of the
product ρ1r3
1(v cos θ)2. This in turn can be used to draw conclusions about
the impactors which made various craters on the solid surfaces of objects in
the planetary system.
5 Acknowledgment
This paper was prepared within the research project 174031 financed by the
Ministry of Education Science and Technological Development of Serbia. I
am grateful to the referee for helpful comments and to Dr. Jean Souchay
from Observatoire de Paris for correspondence .
References
[1] Celebonovic,V.,and Souchay,J.,2010,in
Publ.Astron.Obs.Belgrade,90,205
[2] Celebonovic,V.,2012, Serb.Astron.J.,184,83
[3] Clark, A.H. and Behringer,R.P.,2013, Europhys.Lett.,101,64001
[4] Holsapple, K.A., and Housen, K.R.,2007, Icarus, 187, 345
[5] Dolocan,V.,Dolocan,A.and Dolocan,V.O.,
2008,Mod.Phys.Lett.,B22,2481
[6] King,D. 2007,Guidebook to the Geology of Barringer Meteorite
Crater,Arizona, LPI CONTRIB.1355
7
[7] Landau,L.D.and Lifchitz,E.M. 1976,Statistical Physics,Vol.1.,
§65, Nauka, Moscow (in Russian)
[8] Melosh,H.J. 1989, Impact Cratering,A Geologic Process (Clarendon
Press,Oxford)
[9] Melosh,H.and Collins,G.,2005,Nature,434,157
[10] Pandey,A.K. and Srivastava,A.P. 2011, Int.J.Mod.Phys.,B25, 1543
[11] Ringl,Ch.,Bringa,E.M. and Urbassek,H.M.,2012,Phys.Rev.,E86,061313
[12] Stacey,F.D., 2005,Rep.Progr.Phys.,68,341
[13] Thompson,S.L.1990,ANEOS Analytic Equations of State for Shock
Physics Codes Input Manual,SANDIA REPORT SAND 89-2951
8
|
0905.3633 | 1 | 0905 | 2009-05-22T08:55:34 | CoRoT-2a magnetic activity: hints for possible star-planet interaction | [
"astro-ph.EP",
"astro-ph.SR"
] | CoRoT-2a is a young (about 0.5 Gyr) G7V star accompanied by a transiting hot-Jupiter, discovered by the CoRoT satellite (Alonso et al. 2008; Bouchy et al. 2008). An analysis of its photospheric activity, based on spot modelling techniques previously developed by our group for the analysis of the Sun as a star, shows that the active regions on CoRoT-2a arised within two active longitudes separated by about 180 degrees and rotating with periods of 4.5221 and 4.5543 days, respectively, at epoch of CoRoT observations (112 continous days centered at 2007.6). We show that the total spotted area oscillates with a period of about about 8.9 days, a value close to 10 times the synodic period of the planet with respect to the active longitude pattern rotating in 4.5221 days. Moreover, the variance of the stellar flux is modulated in phase with the planet orbital period. This suggests a possible star-planet magnetic interaction, a phenomenon already seen in other extrasolar planetary systems hosting hot-Jupiters. | astro-ph.EP | astro-ph | CoRoT-2a magnetic activity:
hints for possible star-planet interaction
Isabella Pagano1, Antonino F. Lanza1, Giuseppe Leto1, Sergio Messina1, Pierre
Barge2, Annie Baglin3
1 INAF, Osservatorio Astrofisico di Catania, Italy; 2 Laboratoire
d'Astrophysique de Marseille, France; 3 LESIA, Observatoire de Paris, France
+39-09-7332243
+39-095-2937823
[email protected]
CoRoT-2a is a young (≈0.5 Gyr) G7V star accompanied by a transiting hot-Jupiter, discovered by
the CoRoT satellite (Alonso et al. 2008; Bouchy et al. 2008). An analysis of its photospheric
activity, based on spot modelling techniques previously developed by our group for the analysis of
the Sun as a star, shows that the active regions on CoRoT-2a arised within two active longitudes
separated by about 180 degrees and rotating with periods of 4.5221 and 4.5543 days, respectively,
at epoch of CoRoT observations (112 continous days centered at ≈2007.6). We show that the total
spotted area oscillates with a period of about 28.9 days, a value close to 10 times the synodic
period of the planet with respect to the active longitude pattern rotating in 4.5221 days. Moreover,
the variance of the stellar flux is modulated in phase with the planet orbital period. This suggests a
possible star-planet magnetic interaction, a phenomenon already seen in other extrasolar planetary
systems hosting hot-Jupiters.
planetary systems; hot-Jupiters
SPI: Star Planet Interaction
Introduction
Among the 342 extrasolar planets discovered up to March 2009, about 25% are
massive planets (Mpsini>0.2 MJ) in tight orbits (< 0.1 AU) around their parent
stars. Cuntz et al. (2000) predicted that a giant planet in a close orbit may increase
stellar activity by means of tidal and magnetospheric interactions. Observational
support to such a Stellar-Planet Interaction (SPI) has been provided by several
authors: e.g., Shkolnik et al. (2003, 2005, 2008) found chromospheric emission in
several stars hosting planets to be variable with the planet orbital period; Henry et
al. (2002) and Walker et al. (2008) found hints for photospheric activity
modulation in phase with the planet orbital period.
Several other independent studies have highlighted that a short-period planet can
induce activity in the photosphere and upper atmosphere of its host star. The
1
nature of SPI appears to be strongly affected by both the stellar and planetary
magnetic fields. Shkolnik et al. (2009) and Schmitt (2009) show that for
extrasolar planets around young stars conditions similar to those observed for the
Jupiter-Io system are met and consequently similar processes are expected to
occur on far larger scales. Lanza (2008) describes the SPI considering the
reconnection between the stellar coronal field and the magnetic field of the planet.
Reconnection events produce energetic particles that, moving along magnetic
field lines, impact onto the stellar chromosphere giving rise to a localized hot
spot. Moreover, reconnection events in the corona might affect subphotospheric
dynamo action in those stars producing localized photospheric (and
chromospheric) activity migrating in phase with their planets.
Long baseline and continuous observations provided by CoRoT (Baglin et al.
2006) are optimal to study the photospheric variability of the stars hosting the
planets it detects. CoRoT has recently discovered a hot-Jupiter, CoRoT-2b,
orbiting around a main-sequence G7 star which displays a remarkable
photospheric activity (Alonso et al. 2008; Bouchy et al. 2008). A detailed spot
modelling analysis of the light curve of CoRoT-2a has been performed by Lanza
et al. (2009). Here we focus on the aspects suggesting the presence of a star-planet
magnetic interaction.
Observations
CoRoT-2a was observed from May 16 to October 5, 2007. We extracted from the
data archive the N2 chromatic light curves (Baudin et al. 2006) having a sampling
of 512 s during the first week and 32 s thereinafter. The white colour light curve
shown in Fig. 1 is obtained by the combination of red, green and blue fluxes (see
Lanza et al 2009 for further details on data reduction). Transits and rotational
modulation due to stellar activity are clearly visible.
Results
Active regions derived by spot modelling
The high-quality data allow us to model the evolution of the active regions with a
resolution time of 3.15 days. In Fig. 2 we show the normalized spot filling factor
versus longitude and time, obtained by a maximum entropy spot modelling
2
method (see Lanza et al. 2009 for details). The star shows two active longitudes,
one fixed in the adopted reference frame rotating with a period of 4.5221 days, the
other slowly migrating, which corresponds to a rotation period of 4.5543 days.
Signature of the two active regions can be easily found also in the light curve
folded with the stellar rotation period, i.e. 4.5221 days, when flux is averaged in
phase bins having a width of 0.05, as shown in Fig. 3.
Star-Planet Interaction
In order to search for light curve variability possibly linked to the planet orbit, we
have folded the times series with the planet ephemeris - Pp=1.7429964 d,
HJD0=2454237.5356 after Alonso et al. 2008 - and computed the average flux
and its variance in phase bins of 0.05. A clear wave-like behavior is apparent in
the flux variance vs. orbital phases as shown in Fig. 4. The maximum variability
is observed when the planet is in front of the star, slightly before the planet transit,
and the minimum when it is behind. We have analised the flux variance for the
first 75 days of observations only (i.e., about 43 planet orbits), because the
subsequent subset is useless for our analysis being strongly affected by jitter
instrumental effects that produce outliers (easy to remove) and an increased noise
that pollutes the flux variance.
A further suggestion for SPI comes from the analysis of the total spotted area,
plotted vs. time in Fig. 5 which shows a cyclic oscillation with a period of 28.9 ±
4.8 (1σ) days. Such a period is close to 10 times the synodic period of the planet
with respect to the active longitude pattern rotating in 4.5221 days (Psyn= 2.836d).
Discussion
The suggested SPI can be explained by the model proposed by Lanza (2008).
Specifically, if some subphotospheric dynamo action takes place in the star, the
amplified field emerges once the radial gradient of its intensity reaches some
threshold value. The emergence of the flux might be triggered by a small localized
perturbation of the dynamo effect associated with the planet when it passes by the
most active longitude. An integer number of synodic periods are necessary for the
planet to trigger again field emergence in a given active longitude because the
planetary perturbation can play a role only when the field is already close to the
threshold thanks to the steady amplification provided by the stellar dynamo.
3
Acknowledgments: The present study is based on observations obtained with CoRoT, a space
project operated by the French Space Agency, CNES, with participation of the Science Program of
ESA, ESTEC/RSSD, Austria, Belgium, Brazil, Germany and Spain. AFL, IP, GL and SM have
been partially supported by the Italian Space Agency (ASI) under contract ASI/INAF I/015/07/0,
WPP3170.
References
Alonso, R., et al 2008, A&A 482, L21
Baglin, A., Auvergne, M., Boisnard, L., et al. 2006, 36th COSPAR Scientific Assembly, 36, 3749
Baudin, F., et al., 2006, ESA SP-1306, Editors: M. Fridlund, A. Baglin, J. Lochard and L. Conroy.
ISBN 92-9092-465-9, p.145
Bouchy, F. et al. 2008, A&A 482, L25
Cuntz, M. et al. 2000, ApJ 533, L15
Henry, G.W. et al. 2002, ApJ 577, L111
Lanza, A.F. 2008, A&A 487, 1163
Lanza, A. F., Rodonò, M., Pagano, I., Barge, P., Llebaria, A. 2003, A&A 403, 1135
Lanza, A. F., Pagano, I., Leto, G. et al., 2009, A&A, 493, 193
Schmitt, J.H.M.M., AIP Conf. Proc, Vol 1094, pp. 473, DOI:10.1063/1.3099151
Shkolnik, E. et al. 2003, ApJ 597,1092;
Shkolnik, E. et al. 2005, ApJ 622, 1075
Shkolnik, E. et al. 2008, ApJ 676, 628
Shkolnik, E.,; et al. 2009, AIP Conf. Proc, Vol 1094, 275, DOI:10.1063/1.3099102
Walker, G.A.H. et al. 2008, A&A 482, 691
4
Figure 1. CoRoT-2ab white light time series. Time is measured in HJD since 2,450,000.
5
Figure 2. Isocontours of the ratio f/fmax where f is the spot covering factor and fmax its
maximum value, vs time and longitude (from Lanza et al. 2009). Time is measured in HJD
since 2,450,000.
Figure 3. Flux folded with the stellar rotation period (P=4.5221 days). The minima at phases
≈0.47 and ≈0.85 are due to the two main spotted area whose evolution can be followed in
Fig.1.
6
Figure 4. Flux variance vs planetary orbital phase (in bins of width 0.05). Phases are
reckoned by means of the planet ephemeris P=1.7429964 d, HJD0=2454237.5356 after
Alonso et al. 2008.
Figure 5. Variation of the total spotted area vs time from the ME spot models. The area is
measured in units of the stellar photosphere. The error bars account only for random errors
in the area and have a semi-amplitude of 3 standard deviations.
7
|
1310.2428 | 2 | 1310 | 2013-12-05T10:28:55 | A Super-Jupiter orbiting a late-type star: A refined analysis of microlensing event OGLE-2012-BLG-0406 | [
"astro-ph.EP"
] | We present a detailed analysis of survey and follow-up observations of microlensing event OGLE-2012-BLG-0406 based on data obtained from 10 different observatories. Intensive coverage of the lightcurve, especially the perturbation part, allowed us to accurately measure the parallax effect and lens orbital motion. Combining our measurement of the lens parallax with the angular Einstein radius determined from finite-source effects, we estimate the physical parameters of the lens system. We find that the event was caused by a $2.73\pm 0.43\ M_{\rm J}$ planet orbiting a $0.44\pm 0.07\ M_{\odot}$ early M-type star. The distance to the lens is $4.97\pm 0.29$\ kpc and the projected separation between the host star and its planet at the time of the event is $3.45\pm 0.26$ AU. We find that the additional coverage provided by follow-up observations, especially during the planetary perturbation, leads to a more accurate determination of the physical parameters of the lens. | astro-ph.EP | astro-ph |
DRAFT VERSION OCTOBER 10, 2018
Preprint typeset using LATEX style emulateapj v. 7/15/03
A SUPER-JUPITER ORBITING A LATE-TYPE STAR: A REFINED ANALYSIS OF MICROLENSING EVENT
OGLE-2012-BLG-0406.
Y. TSAPRAS1,R1,♠, J.-Y. CHOIK1, R. A. STREET1,♠, C. HANK1,⋆,‡, V. BOZZAN3,♦, A. GOULDU1,‡, M. DOMINIK∗R3,♦,♠,♣, J.-P.
BEAULIEUP4,♣, A. UDALSKIO1,(cid:4), U. G. JØRGENSENN1,♦, T. SUMIM1,†,
D. M. BRAMICHR2,R6, P. BROWNER3,♦, K. HORNER3,♣, M. HUNDERTMARKR3,♦, S. IPATOVN2, N. KAINSR2,♦, C. SNODGRASSR5,♦, I. A.
AND
STEELER4,
(THE ROBONET COLLABORATION),
K. A. ALSUBAIN2, J. M. ANDERSENN9, S. CALCHI NOVATIN3,N18, Y. DAMERDJIN8, C. DIEHLN15,N17, A. ELYIVN8,N19, E. GIANNININ15, S.
HARDISN1, K. HARPSØEN5, T. C. HINSEN1,N10,N11, D. JUNCHERN1, E. KERINSN12, H. KORHONENN1, C. LIEBIGR3, L. MANCININ13, M.
MATHIASENN1, M. T. PENNYU1, M. RABUSN14, S. RAHVARN6, G. SCARPETTAN3,N4,N16, J. SKOTTFELTN1,N5, J. SOUTHWORTHN7, J.
SURDEJN8, J. TREGLOAN-REEDN7, C. VILELAN7, J. WAMBSGANSSN15,
(THE MINDSTEP COLLABORATION),
J. SKOWRONO1, R. POLESKIO1,U1, S. KOZŁOWSKIO1, Ł. WYRZYKOWSKIO1,O3, M. K. SZYMA ´NSKIO1, M. KUBIAKO1, P.
PIETRUKOWICZO1, G. PIETRZY ´NSKIO1,O2, I. SOSZY ´NSKIO1, K. ULACZYKO1,
(THE OGLE COLLABORATION),
M. D. ALBROWP1, E. BACHELETP2,P3, R. BARRYP11, V. BATISTAP4, A. BHATTACHARYAM10, S. BRILLANTP5, J. A. R. CALDWELLP6, A.
CASSANP4, A. COLEP7, E. CORRALESP4, CH. COUTURESP4, S. DIETERSP2, D. DOMINIS PRESTERP8, J. DONATOWICZP9, P. FOUQUÉP2,P3,
J. GREENHILLP7, S. R. KANEP10, D. KUBASP4,P5, J.-B. MARQUETTEP4, J. MENZIESP12, C. PÈREP4, K. R. POLLARDP1, M. ZUBN15,
G. CHRISTIEU6, D. L. DEPOYU2, S. DONGU3, J. DRUMMONDU9, B. S. GAUDIU1, C. B. HENDERSONU1, K. H. HWANGK1, Y. K. JUNGK1,
A. KAVKAU1, J.-R. KOOU4, C.-U. LEEU4, D. MAOZU10, L. A. G. MONARDU5, T. NATUSCHU6, H. NGANU6, H. PARKK1, R. W. POGGEU1,
(THE PLANET COLLABORATION),
I. PORRITTU7, I.-G. SHINK1, Y. SHVARTZVALDU10, T. G. TANU8, J. C. YEEU1,
(THE µFUN COLLABORATION),
F. ABEM2, D. P. BENNETTM10, I. A. BONDM3, C. S. BOTZLERM4, M. FREEMANM4, A. FUKUIM6, D. FUKUNAGAM2, Y. ITOWM2, N.
KOSHIMOTOM1, C. H. LINGM3, K. MASUDAM2, Y. MATSUBARAM2, Y. MURAKIM2, S. NAMBAM1, K. OHNISHIM7, N. J. RATTENBURYM4,
TO. SAITOM8, D. J. SULLIVANM5, W. L. SWEATMANM3, D. SUZUKIM1, P. J. TRISTRAMM9, N. TSURUMIM2, K. WADAM1, N. YAMAIM11,
P. C. M. YOCKM4 A. YONEHARAM11
(THE MOA COLLABORATION)
1Las Cumbres Observatory Global Telescope Network, 6740 Cortona Drive, suite 102, Goleta, CA 93117, USA
K1Department of Physics, Chungbuk National University, Cheongju 361-763, Republic of Korea
R1School of Physics and Astronomy, Queen Mary University of London, Mile End Road, London E1 4NS, UK
R2European Southern Observatory, Karl-Schwarzschild-Str. 2, 85748 Garching bei München, Germany
R3SUPA, School of Physics & Astronomy, University of St Andrews, North Haugh, St Andrews KY16 9SS, UK
R4Astrophysics Research Institute, Liverpool John Moores University, Liverpool CH41 1LD, UK
R5Max Planck Institute for Solar System Research, Max-Planck-Str. 2, 37191 Katlenburg-Lindau, Germany
R6Qatar Environment and Energy Research Institute, Qatar Foundation, Tornado Tower, Floor 19, P.O. Box 5825, Doha, Qatar
N1Niels Bohr Institute, Astronomical Observatory, Juliane Maries vej 30, 2100 Copenhagen, Denmark
N3Dipartimento di Fisica "E. R. Caianiello", Università di Salerno, Via Giovanni Paolo II n. 132, 84084 Fisciano (SA), Italy
N4International Institute for Advanced Scientific Studies (IIASS), 84019 Vietri sul Mare, (SA), Italy
N5Centre for Star and Planet formation, Geological Museum, Øster Voldgade 5, 1350, Copenhagen, Denmark
N2Qatar Foundation, P.O. Box 5825, Doha, Qatar
N6Dept. of Physics, Sharif University of Technology, P.O. Box 11155-9161, Tehran, Iran
N7Astrophysics Group, Keele University, Staffordshire, ST5 5BG, UK
N8Institut d'Astrophysique et de Géophysique, Allée du 6 Août 17, Sart Tilman, Bât. B5c, 4000 Liége, Belgium
N9Boston University, Astronomy Department, 725 Commonwealth Avenue, Boston, MA 02215, USA
N10Armagh Observatory, College Hill, Armagh, BT61 9DG, Northern Ireland, UK
N11Korea Astronomy and Space Science Institute, 776 Daedukdae-ro, Yuseong-gu, Daejeon 305-348, Korea
N12Jodrell Bank Centre for Astrophysics, University of Manchester, Oxford Road,Manchester, M13 9PL, UK
N13 Max Planck Institute for Astronomy, Königstuhl 17, 69117 Heidelberg, Germany
N16INFN, Gruppo Collegato di Salerno, Sezione di Napoli, Italy
N17Hamburger Sternwarte, Universität Hamburg, Gojenbergsweg 112, 21029 Hamburg, Germany
N18Istituto Internazionale per gli Alti Studi Scientifici (IIASS),84019 Vietri Sul Mare (SA), Italy
N19Main Astronomical Observatory, Academy of Sciences of Ukraine, vul. Akademika Zabolotnoho 27, 03680 Kyiv, Ukraine
O1 Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
O2 Universidad de Concepción, Departamento de Astronomia, Casilla 160-C, Concepción, Chile
O3Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
P1University of Canterbury, Dept. of Physics and Astronomy, Private Bag 4800, 8020 Christchurch, New Zealand
P2Université de Toulouse, UPS-OMP, IRAP, 31400 Toulouse, France
P3CNRS, IRAP, 14 avenue Edouard Belin, 31400 Toulouse, France
P4UPMC-CNRS, UMR7095, Institut d'Astrophysique de Paris, 98bis boulevard Arago, 75014 Paris, France
P5European Southern Observatory (ESO), Alonso de Cordova 3107, Casilla 19001, Santiago 19, Chile
P6McDonald Observatory, 16120 St Hwy Spur 78 #2, Fort Davis, TX 79734, USA
P7School of Math and Physics, University of Tasmania, Private Bag 37, GPO Hobart, 7001 Tasmania, Australia
N14 Instituto de Astrofísica, Facultad de Física, Pontificia Universidad Católica de Chile, Av. Vicuña Mackenna 4860, 7820436 Macul, Santiago, Chile
N15Astronomisches Rechen-Institut, Zentrum für Astronomie der Universität Heidelberg (ZAH), Mönchhofstr. 12-14, 69120 Heidelberg, Germany
2
A REFINED ANALYSIS OF MICROLENSING EVENT OGLE-2012-BLG-0406
P8Physics Department, Faculty of Arts and Sciences, University of Rijeka, Omladinska 14, 51000 Rijeka, Croatia
P9Technical University of Vienna, Department of Computing, Wiedner Hauptstrasse 10, Vienna, Austria
P10Department of Physics & Astronomy, San Francisco State University, 1600 Holloway Avenue, San Francisco, CA 94132, USA
P11Laboratory for Exoplanets and Stellar Astrophysics, Mail Code 667, NASA/GSFC, Bldg 34, Room E317, Greenbelt, MD 20771
P12South African Astronomical Observatory, PO Box 9, Observatory 7935, South Africa
U1Department of Astronomy, Ohio State University, 140 West 18th Avenue, Columbus, OH 43210, USA
U2Department of Physics and Astronomy, Texas A&M University, College Station, TX 77843, USA
U3Kavli Institute for Astronomy and Astrophysics, Peking University, Yi He Yuan Road 5, Hai Dian District, Beijing, 100871, China
U4Korea Astronomy and Space Science Institute, Daejeon 305-348, Republic of Korea
U5Klein Karoo Observatory, Calitzdorp, and Bronberg Observatory, Pretoria, South Africa
U6Auckland Observatory, Auckland, New Zealand
U7Turitea Observatory, Palmerston North, New Zealand
U8Perth Exoplanet Survey Telescope, Perth, Australia
U9Possum Observatory, Patutahi, Gisbourne, New Zealand
U10 School of Physics and Astronomy, Tel-Aviv University, Tel-Aviv 69978, Israel
M1Department of Earth and Space Science, Osaka University, Osaka 560-0043, Japan
M2Solar-Terrestrial Environment Laboratory, Nagoya University, Nagoya, 464-8601, Japan
M3Institute of Information and Mathematical Sciences, Massey University, Private Bag 102-904, North Shore Mail Centre, Auckland, New Zealand
M4Department of Physics, University of Auckland, Private Bag 92-019, Auckland 1001, New Zealand
M5School of Chemical and Physical Sciences, Victoria University, Wellington, New Zealand
M6Okayama Astrophysical Observatory, National Astronomical Observatory of Japan, Asakuchi, Okayama 719-0232, Japan
M7Nagano National College of Technology, Nagano 381-8550, Japan
M8Tokyo Metropolitan College of Aeronautics, Tokyo 116-8523, Japan
M9Mt. John University Observatory, P.O. Box 56, Lake Tekapo 8770, New Zealand
M10University of Notre Dame, Department of Physics, 225 Nieuwland Science Hall, Notre Dame, IN 46556-5670, USA
M11Department of Physics, Faculty of Science, Kyoto Sangyo University, 603-8555, Kyoto, Japan
♠The RoboNet Collaboration
♦The MiNDSTEp Collaboration
(cid:4)The OGLE Collaboration
‡The µFUN Collaboration
♣The PLANET Collaboration
†The MOA Collaboration
and
⋆Corresponding author
Draft version October 10, 2018
ABSTRACT
We present a detailed analysis of survey and follow-up observations of microlensing event OGLE-2012-BLG-
0406 based on data obtained from 10 different observatories. Intensive coverage of the lightcurve, especially
the perturbation part, allowed us to accurately measure the parallax effect and lens orbital motion. Combining
our measurement of the lens parallax with the angular Einstein radius determined from finite-source effects,
we estimate the physical parameters of the lens system. We find that the event was caused by a 2.73 ± 0.43 MJ
planet orbiting a 0.44 ± 0.07 M⊙ early M-type star. The distance to the lens is 4.97 ± 0.29 kpc and the projected
separation between the host star and its planet at the time of the event is 3.45 ± 0.26 AU. We find that the
additional coverage provided by follow-up observations, especially during the planetary perturbation, leads to
a more accurate determination of the physical parameters of the lens.
Subject headings: gravitational lensing -- binaries: general -- planetary systems
1. INTRODUCTION
Radial velocity and transit surveys, which primarily target
main-sequence stars, have already discovered hundreds of gi-
ant planets and are now beginning to explore the reservoir of
lower mass planets with orbit sizes extending to a few as-
tronomical units (AU). These planets mostly lie well inside
the snow line1 of their host stars. Meanwhile, direct imag-
ing with large aperture telescopes has been discovering gi-
ant planets tens to hundreds of AUs away from their stars
(Kalas et al. 2005). The region of sensitivity of microlens-
ing lies somewhere in between and extends to low-mass ex-
oplanets lying beyond the snow-line of their low-mass host
stars, between ∼1 and 10 AU (Tsapras et al. 2003; Gaudi
2012). Although there is already strong evidence that cold
sub-Jovian planets are more common than originally thought
∗Royal Society University Research Fellow
1 The snow line is defined as the distance from the star in a protoplanetary
disk where ice grains can form (Lecar et al. 2006).
around low-mass stars (Gould et al. 2006; Sumi et al. 2010;
Kains et al. 2013; Batalha et al. 2013), cold super-Jupiters or-
biting K or M-dwarfs were believed to be a rarer class of ob-
jects2 (Laughlin et al. 2004; Miguel et al. 2011; Cassan et al.
2012).
Both gravitational instability and core accretion models of
planetary formation have a hard time generating these planets,
although it is possible to produce them given appropriate ini-
tial conditions. The main argument against core accretion is
that it takes too long to produce a massive planet but this cru-
cially depends on the core mass and the opacity of the planet
envelope during gas accretion. In the case of gravitational in-
stability, a massive protoplanetary disc would probably have
too high an opacity to fragment locally at distances of a few
AU.
The radial velocity method has been remarkably successful
2 although a metal-rich protoplanetary disk might allow the formation of
sufficiently massive solid cores.
TSAPRAS ET AL.
in tabulating the part of the distribution that lies within the
snow-line but discoveries of super-Jupiters beyond the snow-
line of M-dwarfs have been comparatively few (Johnson et al.
2010; Montet et al. 2013). Since microlensing is most sensi-
tive to planets that are further away from their host stars, typi-
cally M and K dwarfs, the two techniques are complementary
(Gaudi 2012).
Three brown dwarf and nineteen planet microlensing dis-
coveries have been published to date,
including the dis-
coveries of two multiple-planet systems (Gaudi et al. 2008;
Han et al. 2013)3. It is also worth noting that unbound objects
of planetary mass have also been reported (Sumi et al. 2011).
Microlensing involves the chance alignment along an ob-
server's line of sight of a foreground object (lens) and a back-
ground star (source). This results in a characteristic variation
of the brightness of the background source as it is being grav-
itationally lensed. As seen from the Earth, the brightness of
the source increases as it approaches the lens, reaching a max-
imum value at the time of closest approach. The brightness
then decreases again as the source moves away from the lens.
In microlensing events, planets orbiting the lens star can
reveal their presence through distortions in the otherwise
smoothly varying standard single lens lightcurve. Together,
the host star and planet constitute a binary lens. Binary lenses
have a magnification pattern that is more complex than the
single lens case due to the presence of extended caustics that
represent the positions on the source plane at which the lens-
ing magnification diverges. Distortions in the lightcurve arise
when the trajectory of the source star approaches (or crosses)
the caustics (Mao and Paczy´nski 1991). Recent reviews of the
method can be found in Dominik (2010) and Gaudi (2011).
Upgrades to the OGLE4 (Udalski 2003) survey observing
setup and MOA5 (Sumi et al. 2003) microlensing survey tele-
scope in the past couple of years brought greater precision
and enhanced observing cadence, resulting in an increased
rate of exoplanet discoveries. For example, OGLE has reg-
ularly been monitoring the field of the OGLE-2012-BLG-
0406 event since March 2010 with a cadence of 55 minutes.
When a microlensing alert was issued notifying the astro-
nomical community that event OGLE-2012-BLG-0406 was
exhibiting anomalous behavior, intense follow-up observa-
tions from multiple observatories around the world were ini-
tiated in order to better characterize the deviation. This event
was first analyzed by Poleski et al. (2013) using exclusively
the OGLE-IV survey photometry. That study concluded that
the event was caused by a planetary system consisting of a
3.9±1.2 MJ planet orbiting a low mass late K/early M dwarf.
In this paper we present the analysis of the event based
on the combined data obtained from 10 different telescopes,
spread out in longitude, providing dense and continuous cov-
erage of the lightcurve.
The paper is structured as follows: Details of the discov-
ery of this event, follow-up observations and image analysis
procedures are described in Section 2. Section 3 presents the
methodology of modeling the features of the lightcurve. We
provide a summary and conclude in Section 4.
2. OBSERVATIONS AND DATA
Microlensing event OGLE-2012-BLG-0406 was discov-
ered at equatorial coordinates α = 17h53m18.17s, δ =
3 For a complete list consult http://exoplanet.eu/catalog/ and references
therein.
4 http://ogle.astrouw.edu.pl
5 http://www.phys.canterbury.ac.nz/moa
3
- 30◦28′16.2′′ (J2000.0)6 by the OGLE-IV survey and an-
nounced by their Early Warning System (EWS)7 on the 6th
of April 2012. The event had a baseline I-band magnitude of
16.35 and was gradually increasing in brightness. The pre-
dicted maximum magnification at the time of announcement
was low, therefore the event was considered a low-priority
target for most follow-up teams who preferentially observe
high-magnification events as they are associated with a higher
probability of detecting planets (Griest and Safizadeh 1998).
OGLE observations of the event were carried out with the
1.3-m Warsaw telescope at the Las Campanas Observatory,
Chile, equipped with the 32 chip mosaic camera. The event's
field was visited every 55 minutes providing very dense and
precise coverage of the entire light curve from the baseline,
back to the baseline. For more details on the OGLE data and
coverage see Poleski et al. (2013).
An assessment of data acquired by the OGLE team un-
the 1st of July (08:47 UT, HJD∼2456109.87) which
til
was carried out by the SIGNALMEN anomaly detector
(Dominik et al. 2007) on the 2nd of July (02:19 UT) con-
cluded that a microlensing anomaly, i.e. a deviation from
the standard bell-shaped Paczy´nski curve (Paczy´nski 1986),
was in progress. This was electronically communicated
via the ARTEMiS (Automated Robotic Terrestrial Exoplanet
Microlensing Search) system (Dominik et al. 2008) to trig-
ger prompt observations by both the RoboNet-II8 collabo-
ration (Tsapras et al. 2009) and the MiNDSTEp9 consortium
(Dominik 2010). RoboNet's web-PLOP system (Horne et al.
2009) reacted to the trigger by scheduling observations al-
ready from the 2nd of July (02:30 UT), just 11 minutes af-
ter the SIGNALMEN assessment started. However, the first
RoboNet observations did not occur before the 4th of July
(15:26 UT), when the event was observed with the FTS. This
delayed response was due to the telescopes being offline for
engineering work and bad weather at the observing sites. It
fell to the Danish 1.54m at ESO La Silla to provide the first
data point following the anomaly alert (2nd of July, 03:42 UT)
as part of the MiNDSTEp efforts. The alert also triggered au-
tomated anomaly modeling by RTModel (Bozza 2010), which
by the 2nd of July (04:22 UT) delivered a rather broad variety
of solutions in the stellar binary or planetary range, reflecting
the fact that the true nature was not well-constrained by the
data available at that time. This process chain did not involve
any human interaction at all.
The first human involvement was an e-mail circulated to all
microlensing teams by V. Bozza on the 2nd of July (07:26 UT)
informing the community about the ongoing anomaly and
modeling results. Including OGLE data from a subsequent
night, the apparent anomaly was also independently spotted
by E. Bachelet (e-mail by D.P. Bennett of 3rd July, 13:42 UT),
and subsequently PLANET10 team (Beaulieu et al. 2006)
SAAO data as well as µFUN11 (Gould et al. 2006) SMARTS
(CTIO) data were acquired the coming night, which along
with the RoboNet FTS data cover the main peak of the
anomaly. It should be noted that the observers at CTIO de-
cided to follow the event even while the moon was full in
order to obtain crucial data. A model circulated by T. Sumi
6 (l, b) = - 0.46◦,- 2.22◦
7 http://ogle.astrouw.edu.pl/ogle4/ews/ews.html
8 http://robonet.lcogt.net
9 http://www.mindstep-science.org
10 http://planet.iap.fr
11 http://www.astronomy.ohio-state.edu/∼microfun
4
A REFINED ANALYSIS OF MICROLENSING EVENT OGLE-2012-BLG-0406
TABLE 1. OBSERVATIONS
group
OGLE
RoboNet
RoboNet
RoboNet
MiNDSTEp
MOA
µFUN
PLANET
PLANET
WISE
telescope
passband
data points
1.3m Warsaw Telescope, Las Campanas Observatory (LCO), Chile
2.0m Faulkes North Telescope (FTN), Haleakala, Hawaii, USA
2.0m Faulkes South Telescope (FTS), Siding Spring Observatory (SSO), Australia
2.0m Liverpool Telescope (LT), La Palma, Spain
1.5m Danish Telescope, La Silla, Chile
0.6m Boller & Chivens (B&C), Mt. John, New Zealand
1.3m SMARTS, Cerro Tololo Inter-American Observatory (CTIO), Chile
1.0m Elizabeth Telescope, South African Astronomical Observatory (SAAO), South Africa
1.0m Canopus Telescope, Mt. Canopus Observatory, Tasmania, Australia
1.0m Wise Telescope, Wise Observatory, Israel
I
I
I
I
I
I
V , I
I
I
I
3013
83
121
131
473
1856
16, 81
226
210
180
on the 5th of July (00:38 UT) did not distinguish between the
various solutions.
However, when the rapidly changing features of the
anomaly were independently assessed by the Chungbuk Na-
tional University group (CBNU, C. Han), the community was
informed on the 5th of July (10:43 UT) that the anomaly is
very likely due to the presence of a planetary companion.
An independent modeling run by V. Bozza's automatic soft-
ware (5th of July, 10:55 UT) confirmed the result. While the
OGLE collaboration (A. Udalski) notified observers on the 5th
of July that a caustic exit was occurring, a geometry leading
to a further small peak successively emerged from the mod-
els. D.P. Bennett circulated a model using updated data on
the 6th of July (00:14 UT) which highlighted the presence of
a second prominent feature expected to occur ∼10th of July.
Another modeling run performed at CBNU on the 7th of July
(02:39 UT) also identified this feature and estimated that the
secondary peak would occur on the 11th of July.
Follow-up teams continued to monitor the progress of the
event intensively until the beginning of September, well after
the planetary deviation had ceased, and provided dense cov-
erage of the main peak of the event. A preliminary model
using available OGLE and follow-up data at the time, circu-
lated on the 31st of October (C. Han, J.-Y. Choi), classified the
companion to the lens as a super-Jupiter. Poleski et al. (2013)
presented an analysis of this event using reprocessed survey
data exclusively. In this paper we present a refined analysis
using survey and follow-up data together.
The groups that contributed to the observations of this
event, along with the telescopes used, are listed in Table 1.
Most observations were obtained in the I-band and some im-
ages were also taken in other bands in order to create a color-
magnitude diagram and classify the source star. We note
that there are also observations obtained from the MOA 1.8m
survey telescope which we did not include in our modeling
because the target was very close to the edge of the CCD.
We also do not include data from the µFUN Auckland 0.4m,
PEST 0.3m, Possum 0.36m and Turitea 0.36m telescopes due
to poor observing conditions at site.
Extracting accurate photometry from observations of
crowded fields, such as the Galactic Bulge, is a challenging
process. Each image contains thousands of stars whose stellar
point-spread functions (PSFs) often overlap so aperture and
PSF-fitting photometry can at best offer limited precision. In
order to optimize the photometry it is necessary to use differ-
ence imaging (DI) techniques (Alard and Lupton 1998). For
any particular telescope/camera combination, DI uses a refer-
ence image of the event taken under optimal seeing conditions
which is then degraded to match the seeing conditions of ev-
ery other image of the event taken from that telescope. The
degraded reference image is then subtracted from the match-
ing image to produce a residual (or difference) image. Stars
that have not varied in brightness in the time interval between
the two images will cancel, leaving no systematic residuals
on the difference image but variable stars will leave either a
positive or negative residual.
DI is the preferred method of photometric analysis among
microlensing groups and each group has developed custom
pipelines to reduce their observations. OGLE and MOA im-
ages were reduced using the pipelines described in Udalski
(2003) and Bond et al. (2001) respectively. PLANET, µFUN,
and WISE images were processed using variants of the PySIS
(Albrow et al. 2009) pipeline, whereas RoboNet and MiND-
STEp observations were analyzed using customized versions
of the DanDIA package (Bramich 2008). Once the source star
returned to its baseline magnitude, each data set was repro-
cessed to optimize photometric precision. These photometri-
cally optimized data sets were used as input for our modeling
run.
3. MODELING
Figure 1 shows the lightcurve of OGLE-2012-BLG-406.
The lightcurve displays two main features that deviate sig-
nificantly from the standard Paczy´nski curve. The first fea-
ture, which peaked at HJD ∼ 2456112 (3rd of July), is pro-
duced by the source trajectory grazing the cusp of a caustic.
The brightness then quickly drops as the source moves away
from the cusp (Schneider & Weiss 1992; Zakharov 1995), in-
creases again for a brief period as it passes close to another
cusp at HJD ∼ 2456121 (12th of July), and eventually returns
to the standard shape as the source moves further away from
the caustic structure. The anomalous behavior, when both fea-
tures are considered, lasts for a total of ∼ 15 days, while the
full duration of the event is &120 days. These are typical
lightcurve features expected from lensing phenomena involv-
ing planetary lenses.
We begin our analysis by exploring a standard set of solu-
tions that involve modeling the event as a static binary lens.
The Paczy´nski curve representing the evolution of the event
for most of its duration is described by three parameters: the
time of closest approach between the projected position of the
source on the lens plane and the position of the lens photo-
center12, t0, the minimum impact parameter of the source, u0,
12 The "photocenter" refers to the center of the lensing magnification pat-
tern. For a binary-lens with a projected separation between the lens compo-
nents less than the Einstein radius of the lens, the photocenter corresponds
to the center of mass. For a lens with a separation greater than the Einstein
radius, there exist two photocenters each of which is located close to each
TSAPRAS ET AL.
5
FIG. 1. -- Lightcurve of OGLE-2012-BLG-0406 showing our best-fit binary-lens model including parallax and orbital motion. The legend on the right of the
figure lists the contributing telescopes. All data were taken in the I-band, except where otherwise indicated.
parameters
standard
parallax
u0 > 0
u0 < 0
u0 > 0
orbit
u0 < 0
TABLE 2. LENSING PARAMETERS
orbit+parallax
u0 > 0
u0 < 0
χ2/dof
t0 (HJD')
u0
tE (days)
s
q (10- 3)
α
ρ∗ (10- 2)
πE,N
πE,E
ds/dt (yr- 1)
dα/dt (yr- 1)
6921.019/6383
6141.63 ± 0.04
0.532 ± 0.001
62.37 ± 0.06
1.346 ± 0.001
5.33 ± 0.04
0.852 ± 0.001
1.103 ± 0.008
--
--
--
--
6850.358/6381
6141.70 ± 0.05
0.527 ± 0.001
63.75 ± 0.18
1.345 ± 0.001
5.07 ± 0.03
0.864 ± 0.002
1.053 ± 0.007
0.118 ± 0.011
-0.033 ± 0.007
--
--
6677.685/6381
6141.66 ± 0.05
-0.520 ± 0.001
69.39 ± 0.32
1.341 ± 0.001
4.45 ± 0.04
-0.906 ± 0.002
0.968 ± 0.009
-0.414 ± 0.016
-0.069 ± 0.009
--
--
6408.371/6381
6141.24 ± 0.05
0.500 ± 0.002
65.33 ± 0.20
1.300 ± 0.002
6.97 ± 0.27
0.861 ± 0.002
1.233 ± 0.031
--
--
0.765 ± 0.046
1.284 ± 0.159
6408.255/6381
6141.28 ± 0.04
-0.499 ± 0.002
65.53 ± 0.15
1.301 ± 0.001
6.63 ± 0.05
-0.859 ± 0.001
1.194 ± 0.011
--
--
0.727 ± 0.017
-1.108 ± 0.019
6357.680/6379
6141.33 ± 0.05
0.496 ± 0.002
64.77 ± 0.19
1.301 ± 0.002
5.92 ± 0.11
0.837 ± 0.002
1.111 ± 0.014
-0.143 ± 0.018
0.047 ± 0.007
0.669 ± 0.028
0.497 ± 0.059
6381.358/6379
6141.19 ± 0.06
-0.497 ± 0.002
61.91 ± 0.42
1.296 ± 0.002
6.82 ± 0.19
-0.810 ± 0.005
1.207 ± 0.023
0.358 ± 0.042
0.008 ± 0.006
0.802 ± 0.033
-0.732 ± 0.085
NOTE. -- HJD'=HJD-2450000.
expressed in units of the angular Einstein radius of the lens
(θE), and the duration of time, tE (the Einstein time-scale), re-
quired for the source to cross θE. The binary nature of the lens
requires the introduction of three extra parameters: The mass
ratio q between the two components of the lens, their pro-
jected separation s, expressed in units of θE, and the source
trajectory angle α with respect to the axis defined by the two
components of the lens. A seventh parameter, ρ∗, representing
the source radius normalized by the angular Einstein radius
is also required to account for finite-source effects that are
lens component with an offset q/[s(1 + q)] toward the other lens component
(Kim et al. 2009). In this case, the reference t0, u0 measurement is obtained
from the photocenter to which the source trajectory approaches closest.
important when the source trajectory approaches or crosses a
caustic (Ingrosso et al. 2009).
The magnification pattern produced by binary lenses is very
sensitive to variations in s,q, which are the parameters that
affect the shape and orientation of the caustics, and α, the
source trajectory angle. Even small changes in these param-
eters can produce extreme changes in magnification as they
may result in the trajectory of the source approaching or cross-
ing a caustic (Dong et al. 2006, 2009). On the other hand,
changes in the other parameters cause the overall magnifica-
tion pattern to vary smoothly.
To assess how the magnification pattern depends on the pa-
rameters, we start the modeling run by performing a hybrid
6
A REFINED ANALYSIS OF MICROLENSING EVENT OGLE-2012-BLG-0406
search in parameter space whereby we explore a grid of s,q, α
values and optimize t0,u0,tE and ρ∗ at each grid point by
χ2 minimization using Markov Chain Monte Carlo (MCMC).
Our grid limits are set at - 1 ≤ logs ≤ 1, - 5 ≤ log q ≤ 1, and
0 ≤ α < 2π, which are wide enough to guarantee that all local
minima in parameter space have been identified. An initial
MCMC run provides a map of the topology of the χ2 surface,
which is subsequently further refined by gradually narrow-
ing down the grid parameter search space (Shin et al. 2012a;
Street et al. 2013). Once we know the approximate locations
of the local minima, we perform a χ2 optimization using all
seven parameters at each of those locations in order to deter-
mine the refined position of the minimum. From this set of
local minima, we identify the location of the global minimum
and check for the possible existence of degenerate solutions.
We find no other solutions.
FIG. 2. -- The bottom panel zooms-in on the anomalous region of the
lightcurve presented in Figure 1. At the top panel we display the source tra-
jectory, color coded for the individual contributions of each observatory, and
caustic structure at two different times corresponding to the first and sec-
ond peaks of the anomaly. All scales are normalized by θE, and the size of
the circles corresponds to the size of the source. The first peak deviates the
strongest. This is a result of the trajectory of the source grazing the cusp of
the caustic at t1 (HJD∼2456112), shown in red. The second deviation at t2
(HJD∼2456121) is significantly weaker and is due to the source trajectory
passing close to another cusp of the caustic, shown in blue. The differences
in the shape of the caustic shown at t1 and t2 are due to the orbital motion of
the lens-planet system.
Since our analysis relies on data sets obtained from dif-
ferent telescopes and instruments which use different esti-
mates for the reported photometric precision, we normalize
the flux uncertainties of each data set by adjusting them as
+ σ2
ei = fi(σ2
i )1/2, where fi is a scale factor, σ0 are the orig-
0
inally reported uncertainties and σi is an additive uncertainty
term for each data set i. The rescaling ensures that χ2 per de-
gree of freedom (χ2/dof) for each data set relative to the model
becomes unity. Data points with very large uncertainties and
obvious outliers are also removed in the process.
In computing finite-source magnifications, we take into ac-
count the limb-darkening of the source by modeling the sur-
face brightness as Sλ(ϑ) ∝ 1 - Γλ(1 - 1.5 cosϑ) (Albrow et al.
2001), where ϑ is the angle between the line of sight toward
the source star and the normal to the source surface, and Γλ
is the limb-darkening coefficient in passband λ. We adopt
ΓV = 0.74 and ΓI = 0.53 from the Claret (2000) tables. These
values are based on our classification of the stellar type of the
source, as subsequently described.
The residuals contained additional smooth structure that the
static binary model did not account for. This indicated the
need to consider additional second-order effects. The event
lasted for & 120 days, so the positional change of the ob-
server caused by the orbital motion of the Earth around the
Sun may have affected the lensing magnification. This intro-
duces subtle long-term perturbations in the event lightcurve
by causing the apparent lens-source motion to deviate from a
rectilinear trajectory (Gould 1992; Alcock et al. 1995). Mod-
eling this parallax effect requires the introduction of two ex-
tra parameters, πE,N and πE,E, representing the components of
the parallax vector πE projected on the sky along the north
and east equatorial axes respectively. When parallax effects
are included in the model, we use the geocentric formalism
of Gould et al. (2004) which ensures that the parameters t0,
u0 and tE will be almost the same as when the event is fitted
without parallax.
An additional effect that needs to be considered is the or-
bital motion of the lens system. The lens orbital motion
causes the shape of the caustics to vary with time. To a first or-
der approximation, the orbital effect can be modeled by intro-
ducing two extra parameters that represent the rate of change
of the normalized separation between the two lensing com-
ponents ds/dt and the rate of change of the source trajectory
angle relative to the caustics dα/dt (Albrow et al. 2000).
We conduct further modeling considering each of the
higher-order effects separately and also model their combined
effect. Furthermore, for each run considering a higher-order
effect, we test models with u0 > 0 and u0 < 0 that form a pair
of degenerate solutions resulting from the mirror-image sym-
metry of the source trajectory with respect to the binary-lens
axis. For each model, we repeat our calculations starting from
different initial positions in parameter space to verify that the
fits converge to our previous solution and that there are no
other possible minima.
Table 2 lists the optimized parameters for the models we
considered. We find that higher-order effects contribute
strongly to the shape of the lightcurve. The model includ-
ing the parallax effect provides a better fit than the standard
model by ∆χ2 = 243.3. The orbital effect also improves the fit
by ∆χ2 = 512.8. The combination of both parallax and orbital
effects improves the fit by ∆χ2 = 563.3. Due to the u0 > 0 and
u0 < 0 degeneracy, there are two solutions for the orbital mo-
tion + parallax model which have similar χ2 values. Models
involving the xallarap effect (source orbital motion) were also
considered but they did not outperform equivalent models in-
volving only parallax.
In Figure 1, we present the best-fit model lightcurve super-
posed on the observed data. Figure 2 displays an enlarged
TSAPRAS ET AL.
7
view of the perturbation region of the lightcurve along with
the source trajectory with respect to the caustic. The follow-
up observations cover critical features of the perturbation re-
gions that were not covered by the survey data. We note that
the caustic varies with time and thus we present the shape of
the caustic at the times of the first (t1=HJD∼2456112) and
second perturbations (t2=HJD∼2456121). The source trajec-
tory grazes the caustic structure at t1 causing a substantial in-
crease in magnification. As the caustic structure and trajec-
tory evolve with time, the trajectory approaches another cusp
at t2, but does not cross it. This second approach causes an in-
crease in magnification which is appreciably lower than that
of the first encounter at t1. The source trajectory is curved due
to the combination of the parallax and orbital effects.
The mass and distance to the lens are determined by
Mtot =
θE
κπE
;
DL =
AU
πEθE + πS
,
(1)
where κ = 4G/(c2AU) and πS is the parallax of the source
star (Gould 1992). To determine these physical quantities we
require the values of πE and θE. Modeling the event returns
the value of πE, whereas θE = θ∗/ρ∗ depends on the angular
radius of the source star, θ∗, and the normalized source ra-
dius, ρ∗, which is also returned from modeling (see Table 2).
Therefore, determining θE requires an estimate of θ∗.
To estimate the angular source radius, we use the stan-
dard method described in Yoo et al. (2004).
In this proce-
dure we first measure the dereddened color and brightness
of the source star by using the centroid of the giant clump
as a reference because its dereddened magnitude I0,c = 14.45
(Nataf et al. 2013) and color (V -
I)0,c = 1.06 (Bensby et al.
2011) are already known. For this calibration, we use a color-
magnitude diagram obtained from CTIO observations in the
I and V bands. We then convert the V -
I source color to
V - K using the color-color relations from Bessell and Brett
(1988) and the source radius is obtained from the θ∗-(V - K)
relations of Kervella et al. (2004). We derive the dereddened
magnitude and color of the source star as I0 = 14.62 and
(V -
I)0 = 1.12 respectively. This confirms that the source star
is an early K-type giant. The estimated angular source radius
is θ∗ = 5.94 ± 0.51 µas. Combining this with our evaluation
of ρ∗, we obtain θE = 0.53 ± 0.05 mas for the angular Einstein
radius of the lens.
Our analysis is consistent with the results of Poleski et al.
(2013). We confirm that the lens is a planetary system com-
posed of a giant planet orbiting a low-mass star and we report
the refined parameters of the system. Poleski et al. (2013) re-
ported that there existed a pair of degenerate solutions with
u0 > 0 and u0 < 0, although the positive u0 solution is slightly
preferred with ∆χ2 = 13.6. We find a consistent result that
the positive u0 solution is preferred but the degeneracy is bet-
ter discriminated by ∆χ2 = 23.7.
The error contours of the parallax parameters for the best-fit
model are presented in Figure 3. The uncertainty of each pa-
rameter is determined from the distribution of MCMC chain,
and the reported uncertainty corresponds to the standard devi-
ation of the distribution. We list the physical parameters of the
system in Table 3 and their posterior probability distributions
are shown in Figure 4.
The lens lies DL = 4.97 ± 0.29 kpc away in the direction
of the Galactic Bulge. The more massive component of the
lens has mass M⋆ = 0.44 ± 0.07 M⊙ so it is an early M-type
dwarf star and its companion is a super-Jupiter planet with a
mass Mp = 2.73 ± 0.43 MJ. The projected separation between
FIG. 3. -- ∆χ2 contours for the parallax parameters derived from our
MCMC fits for the best binary-lens model including orbital motion and the
parallax effect.
TABLE 3. PHYSICAL PARAMETERS
parameters
quantity
Mass of the host star (M⋆)
Mass of the planet (Mp)
Distance to the lens (DL)
Projected star-planet separation (d⊥)
Einstein radius (θE)
Geocentric proper motion (µGeo)
0.44 ± 0.07 M⊙
2.73 ± 0.43 MJ
4.97 ± 0.29 kpc
3.45 ± 0.26 AU
0.53 ± 0.05 milli-arcsec
3.02 ± 0.26 milli-arcsec yr- 1
the two components of the lens is d⊥ = 3.45 ± 0.26 AU. The
geocentric relative proper motion between the lens and the
source is µGeo = θE/tE = 3.02 ± 0.26 milli-arcsec yr- 1. In the
Heliocentric frame, the proper motion is µHelio = (µN, µE) =
(- 2.91 ± 0.26,1.31 ± 0.16) milli-arcsec yr- 1.
We note that the derived physical lens parameters are some-
what different from those of Poleski et al. (2013). Specifi-
cally, the mass of the host star derived in Poleski et al. (2013)
is 0.59 M⊙, which is ∼ 34% greater than our estimate. Half
of this difference comes from the slightly larger Einstein ra-
dius obtained by Poleski et al. (2013) from the OGLE-IV pho-
tometry and the remaining part from the slightly larger πE,N
component of the parallax obtained from modeling the sur-
vey and follow-up photometry as presented in this paper. It
should be noted that the parameters derived by both our and
thePoleski et al. (2013) models are consistent within the 1-σ
level.
To further check the consistency between our model and
that of Poleski et al. (2013), we conducted additional mod-
eling based on different combinations of data sets. We first
test a model based on OGLE data exclusively in order to see
whether we can retrieve the physical parameters reported in
Poleski et al. (2013). From this modeling, we derive physi-
cal parameters consistent with those of Poleski et al. (2013),
indicating that the differences are due to the additional cov-
8
A REFINED ANALYSIS OF MICROLENSING EVENT OGLE-2012-BLG-0406
FIG. 4. -- The physical parameter uncertainties pertaining to the lens as derived from the MCMC runs optimizing our binary-lens model including parallax and
orbital motion for the u0 > 0 trajectory.
mass star discovered by microlensing (Dong et al. 2009;
Batista et al. 2011; Yee et al. 2012) and the first such system
whose characteristics were derived solely from microlensing
data, without considering any external information.
erage provided by the follow-up observations. We conducted
another modeling run using OGLE observations but also in-
cluded CTIO, FTS and SAAO data, i.e. those datasets cover-
ing the anomalous peak. This modeling run resulted in phys-
ical parameters that are consistent with the values extracted
from fitting all combined data together, as reported in this pa-
per. This indicates that the differences between Poleski et al.
(2013) and this analysis, although consistent within the 1-σ
level, come mainly from follow-up data that provide better
coverage of the perturbation. Therefore, using survey and
follow-up data together, we arrive at a more accurate deter-
mination of the ρ and πE,N parameters, which leads to a re-
finement of the physical parameters of the planetary system.
4. CONCLUSIONS
Microlensing event OGLE-2012-BLG-0406 was inten-
sively observed by survey and follow-up groups using 10 dif-
ferent telescopes around the world. Anomalous deviations ob-
served in the lightcurve were recognized to be due to the pres-
ence of a planetary companion even before the event reached
its central peak. The anomalous behavior was first identified
and assessed automatically via software agents. Most follow-
up teams responded to these alerts by adjusting their observ-
ing strategies accordingly. This highlights the importance of
circulating early models to the astronomical community that
help to identify important targets for follow-up observations
(Shin et al. 2012b). There are ∼100 follow-up alerts circu-
lated annually, ∼10% of which turn out to be planet candi-
dates.
Our analysis of the combined data is consistent with the
results of Poleski et al. (2013) and we report the refined pa-
rameters of the system. We find that this refinement is mainly
due to follow-up observations over the anomaly. The primary
lens with mass M⋆ = 0.44 ± 0.07 M⊙ is orbited by a planetary
companion with mass Mp = 2.73 ± 0.43 MJ at a projected sep-
aration of d⊥ = 3.45 ± 0.26 AU. The distance to the system is
DL = 4.97 ± 0.29 kpc in the direction of the Galactic Bulge.
This is the fourth cold super-Jupiter planet around a low-
Microlensing is currently the only way to obtain high pre-
cision mass measurements for this type of system. Radial ve-
locity, in addition to the msin i degeneracy, at present does
not have long enough data streams to measure the parame-
ters of such systems. However, recently Montet et al. (2013)
have developed a promising new method to discover them us-
ing a combination of radial velocity and direct imaging. They
identify long term trends in radial velocity data and use adap-
tive optics imaging to rule out the possibility that these are
due to stars. This means that the trends are either due to
large planets or brown dwarfs. This approach does not yield
precise characterization but provides important statistical in-
formation. Their results are consistent with gravitational mi-
crolensing estimates of planet abundance in that region of pa-
rameter space.
The precise mechanism of how such large planets form and
evolve around low mass stars is still an open question. Ra-
dial velocity and transit surveys have been finding massive
gas-giant planets around FGK-stars for years (Batalha et al.
2013) but these stars have protoplanetary disks that are suffi-
ciently massive to allow the formation of super-Jupiter plan-
ets. On the other hand, protoplanetary disks around M-dwarfs
have masses of only a few Jupiter mass so massive gas giants
should be relatively hard to produce (Apai 2013).
Recent observational studies have revealed that protoplan-
etary disks are as common around low mass stars as higher-
mass stars (Williams & Cieza 2011), arguing for the same for-
mation processes. In addition, there is mounting evidence, but
not yet conclusive, that disks last much longer around low-
mass stars (Apai 2013). Longer disk lifetimes may be con-
ducive to the formation of super-Jupiters. The microlensing
discoveries suggest that giant planets around low-mass stars
may be as common as around higher-mass stars but may not
undergo significant migration (Gould et al. 2010).
TSAPRAS ET AL.
9
Simulations using the core accretion formalism can produce
such planets within reasonable disk lifetimes of a few Myr
(Mordasini et al. 2012) provided the core mass is sufficiently
large or the opacity of the planet envelope during gas accre-
tion is decreased by assuming that the dust grains have grown
to larger sizes than the typical interstellar values (R. Nelson,
private communication). Furthermore, gravitational instabil-
ity models of planet formation can also potentially produce
such objects when the opacity of the protoplanetary disk is
low enough to allow local fragmentation at greater distances
from the host star, and subsequently migrating the planet to
distances of a few AU.
It is worth noting that highly magnified microlensing events
involving extended stellar sources may produce appreciable
polarization signals (Ingrosso et al. 2012). If such signals are
observed during a microlensing event, they can be combined
with photometric observations to place further constraints on
the lensing geometry and physical properties of the lens.
YT thanks the CBNU group for their advice and hospi-
tality while in Korea. DMB, MD, KH, CS, RAS, KAA,
MH and YT are supported by NPRP grant NPRP-09-476-
1-78 from the Qatar National Research Fund (a member of
Qatar Foundation). CS received funding from the European
Union Seventh Framework Programme (FP7/2007-2013) un-
der grant agreement no. 268421. KH is supported by a
Royal Society Leverhulme Trust Senior Research Fellow-
ship. JPB and PF acknowledge the financial support of Pro-
gramme National de PlanÃl'tologie and of IAP. The OGLE
project has received funding from the European Research
Council under the European Community's Seventh Frame-
work Programme (FP7/2007-2013) / ERC grant agreement
no. 246678 to AU. Work by CH was supported by Cre-
ative Research Initiative Program (2009-0081561) of Na-
tional Research Foundation of Korea. The MOA experiment
was supported by grants JSPS22403003 and JSPS23340064.
TS acknowledges the support JSPS24253004. TS is sup-
ported by the grant JSPS23340044. TCH acknowledges
support from KRCF via the KRCF Young Scientist Fellow-
ship program and financial support from KASI grant num-
ber 2013-9-400-00. YM acknowledges support from JSPS
grants JSPS23540339 and JSPS19340058. AG and BSG ac-
knowledge support from NSF AST-1103471. MR acknowl-
edges support from FONDECYT postdoctoral fellowship
No3120097. BSG, AG, and RWP acknowledge support from
NASA grant NNX12AB99G. YD, AE and JS acknowledge
support from the Communaute francaise de Belgique - Ac-
tions de recherche concertees - Academie Wallonie-Europe.
This work is based in part on data collected by MiNDSTEp
with the Danish 1.54m telescope at the ESO La Silla Obser-
vatory. The Danish 1.54m telescope is operated based on a
grant from the Danish Natural Science Foundation (FNU).
REFERENCES
Alard, C. & Lupton, R. H. 1998, ApJ, 503, 325
Albrow, M. D., An, J., Beaulieu, J.-P., et al. 2001, ApJ, 549, 759
Albrow, M. D., Beaulieu, J.-P., Caldwell, J. A. R., et al. 2000, ApJ, 534, 894
Albrow, M. D., Horne, K., Bramich, D. M., et al. 2009, MNRAS, 397, 2099
Alcock, C., Allsman, R. A., Alves, D., et al. 1995, ApJ, 454, 125
Apai, D., 2013, Astronomische Nachrichten, 334, 57
Batalha, N. M., Rowe, J. F., Bryson, S. T., et al. 2013, ApJS, 204, 24
Batista, V., Gould, A., Dieters, S., et al. 2011, A&A, 529, 102
Beaulieu, J.-P., Bennett, D. P., Fouqué, P., et al. 2006, Nature, 439, 437
Bensby, T., Adén, D., Meléndez, J., et al. 2011, A&A, 533, A134
Bessell, M. S. & Brett, J. M. 1988, PASP, 100, 1134
Bond, I. A., Abe, F., Dodd, R. J., et al. 2001, MNRAS, 327, 868
Bozza, V. 2010, MNRAS, 408, 2188
Bramich, D. M. 2008, MNRAS, 386, L77
Cassan, A., Kubas, D., Beaulieu, J.-P., et al. 2012, Nature, 481, 167
Claret, A. 2000, A&A, 363, 1081
Kim, D., & Han, C., & Park, B.-J. 2009, JKAS, 42, 39
Dominik, M. 2010, Generay Relativity and Gravitation, 42, 9
Dominik, M., Horne, K., Allan, A., et al. 2008, Astronomische Nachrichten,
Dominik, M., Jørgensen, U. G., Rattenbury, N. J., et al. 2010, Astronomische
329, 248
Nachrichten, 331, 671
Dominik, M., Rattenbury, N.J., Allan, A., et al. 2007, MNRAS, 380, 792
Dong, S., Bond, I. A., Gould, A., et al. 2009, ApJ, 698, 1826
Dong, S., DePoy, D. L., Gaudi, B. S., et al. 2006, ApJ, 642, 842
Dong, S., Gould, A., Udalski, A., et al. 2009, ApJ, 695, 970
Gaudi, B. S. 2011, Exoplanets (book), 79
Gaudi, B. S. 2012, ARA&A, 50, 411
Gaudi, B. S., Bennett, D. P., Udalski, A., et al. 2008, Science, 319, 927
Gould, A. 1992, ApJ, 392, 442
Gould, A. 2004, ApJ, 606, 319
Gould, A., Udalski, A., An, D., et al. 2006, ApJ, 644, L37
Gould, A., Dong, S., Gaudi, B. S., et al 2010, ApJ, 720, 1073
Griest, K., & Safizadeh, N. 1998, ApJ, 500, 37
Han, C., Udalski, A., Choi, J.-Y., et al. 2013, ApJ, 762, 28
Horne, K., Snodgrass, C., Tsapras, Y., 2009, MNRAS, 396, 2087
Ingrosso, G., Novati, S. Calchi, de Paolis, F., et al. 2012, MNRAS, 426, 1496
Ingrosso, G., Novati, S. Calchi, de Paolis, F., et al. 2009, MNRAS, 399, 219
Johnson, J. A., Howard, A. W., Marcy, G. W., et al. 2010, PASP, 122, 149
Kains, N., Street, R. A., Choi, J.-Y., et al. 2013, A&A, 552, 70
Kalas, P. Graham, J. R., Clampin, M. 2005, Nature, 435, 1067
Kervella, P., Bersier, D. Mourard, D. et al. 2004, A&A, 428, 587
Laughlin, G., Bodenheimer, P., Adams, F. C. 2004, ApJ, 612, 73
Lecar, M., Podolak, M., Sasselov, D., Chiang, E. 2006, ApJ, 640, 1115
Mao, S., & Paczy´nski, B. 1991, ApJ, 374, L37
Miguel, Y., Guilera, O. M., Brunini, A., 2011, MNRAS, 417, 314
Montet, B. T., Crepp, J. R., Johnson, J. A., et al. 2013, ApJ, submitted
Mordasini, C., Alibert, Y., Benz, W., et al. 2012, A&A, 547, 112
Nataf, D. M., Gould, A., Fouqué, P., et al. 2013, ApJ, 769, 88
Paczy´nski, B. 1986, ApJ, 304, 1
Poleski, R., Udalski, A., Dong, S., et al. 2013, ApJ, submitted
Schneider, P. & Weiss, A. 1992, A&A, 260, 1
Shin, I.-G., Choi, J.-Y., Park, S.-Y., et al. 2012, ApJ, 746, 127
Shin, I.-G., Han, C., Gould, A., et al. 2012, ApJ, 760, 116
Street, R., Choi, J.-Y., Tsapras, Y., et al. 2013, ApJ, 763, 67
Sumi, T., Abe, F., Bond, I. A., et al. 2003, ApJ, 591, 204
Sumi, T., Bennett, D. P., Bond, I. A., et al. 2010, ApJ, 710, 1641
Sumi, T., Kamiya, K., Bennett, D. P., et al. 2011, Nature, 473, 349
Tsapras, Y., Horne, K., Kane, S., et al. 2003, MNRAS, 343, 1131
Tsapras, Y., Street, R., Horne, K., et al. 2009, Astronomische Nachrichten,
330, 4
Udalski, A. 2003, Acta Astron., 53, 291
Williams, P. W., Cieza, A. L., 2011, ARA&A, 49, 67
Yee, J., Shvartzvald, Y., Gal-Yam, A., et al. 2012, ApJ, 775, 102
Yoo, J., DePoy, D. L., Gal-Yam, A., et al. 2004, ApJ, 603, 139
Zakharov, A. F. 1995, A&A, 293, 1
|
1810.05418 | 2 | 1810 | 2018-12-10T00:18:12 | Nightside Winds at the Lower Clouds of Venus with Akatsuki/IR2: Longitudinal, local time and decadal variations from comparison with previous measurements | [
"astro-ph.EP",
"physics.ao-ph"
] | We present measurements of the wind speeds at the nightside lower clouds of Venus from observations by JAXA's mission Akatsuki during 2016, complemented with new wind measurements from ground-based observations acquired with TNG/NICS in 2012 and IRTF/SpeX in 2015 and 2017. Zonal and meridional components of the winds were measured from cloud tracking on a total of 466 Akatsuki images of Venus acquired by the camera IR2 using the 2.26-$\mathrm{\mu m}$ filter, with spatial resolutions ranging 10--80 km per pixel and covering from 2016 March 22 to October 31. More than 149,000 wind vectors were obtained with an automatic technique of template matching, and 2,947 wind vectors were inferred with the manual procedure. The meridional profiles for both components of the winds are found to be consistent with results from the Venus Express mission during 2006--2008, although stronger wind variability is found for the zonal component at equatorial latitudes where Akatsuki observations have better viewing geometry than Venus Express. The zonal winds at low latitudes also suggest a zonal variability that could be associated with solar tides or vertically propagating orographic waves. Finally, the combination of our wind measurements from TNG/NICS, IRTF/SpeX and Akatsuki images with previously published and based in data from 1978 to 2017 suggests variations of up to 30 m s$^{-1}$ in the winds at the lower clouds of the Venus nightside. | astro-ph.EP | astro-ph | Draft version December 11, 2018
Typeset using LATEX twocolumn style in AASTeX61
8
1
0
2
c
e
D
0
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
8
1
4
5
0
.
0
1
8
1
:
v
i
X
r
a
NIGHTSIDE WINDS AT THE LOWER CLOUDS OF VENUS WITH AKATSUKI/IR2: LONGITUDINAL,
LOCAL TIME AND DECADAL VARIATIONS FROM COMPARISON WITH PREVIOUS MEASUREMENTS.
Javier Peralta,1 Keishiro Muto,2 Ricardo Hueso,3 Takeshi Horinouchi,4 Agust´ın S´anchez-Lavega,3
Shin-ya Murakami,1 Pedro Machado,5 Eliot F. Young,6 Yeon Joo Lee,2 Toru Kouyama,7 Hideo Sagawa,8
Kevin McGouldrick,9 Takehiko Satoh,1, 10 Takeshi Imamura,2 Sanjay S. Limaye,11 Takao M. Sato,12, 1
Kazunori Ogohara,13 Masato Nakamura,1 and David Luz5
1Institute of Space and Astronautical Science (ISAS), Japan Aerospace Exploration Agency (JAXA)
3-1-1, Yoshinodai, Chuo-ku, Sagamihara, Kanagawa, 252-5210, Japan
2Graduate School of Frontier Sciences, The University of Tokyo, Japan
3Escuela de Ingenier´ıa de Bilbao (UPV/EHU), Bilbao, Spain
4Faculty of Environmental Earth Science, Hokkaido University, Sapporo, Japan
5Institute of Astrophysics and Space Sciences, Portugal
6Southwest Research Institute, Boulder, CO 80302, USA
7Artificial Intelligence Research Center, National Institute of Advanced Industrial Science and Technology, Tokyo, Japan
8Faculty of Science, Kyoto Sangyo University, Japan
9Laboratory for Atmospheric and Space Physics, University of Colorado Boulder, Boulder, CO 80303-7814, USA
10Department of Space and Astronautical Science, School of Physical Sciences, Sokendai, Japan
11Space Science and Engineering Center, University of Wisconsin-Madison, Madison, USA
12Space Information Center, Hokkaido Information University, Hokkaido, Japan
13School of Engineering, University of Shiga Prefecture, Shiga, Japan
(Dated: Published in Astrophysical Journal Supplement Series the 7 of December 2018; original manuscript here.)
ABSTRACT
We present measurements of the wind speeds at the nightside lower clouds of Venus from observations by JAXA's
mission Akatsuki during 2016, complemented with new wind measurements from ground-based observations acquired
with TNG/NICS in 2012 and IRTF/SpeX in 2015 and 2017. Zonal and meridional components of the winds were
measured from cloud tracking on a total of 466 Akatsuki images of Venus acquired by the camera IR2 using the 2.26-
µm filter, with spatial resolutions ranging 10 -- 80 km per pixel and covering from 2016 March 22 to October 31. More
than 149,000 wind vectors were obtained with an automatic technique of template matching, and 2,947 wind vectors
were inferred with the manual procedure. The meridional profiles for both components of the winds are found to be
consistent with results from the Venus Express mission during 2006 -- 2008, although stronger wind variability is found
for the zonal component at equatorial latitudes where Akatsuki observations have better viewing geometry than Venus
Express. The zonal winds at low latitudes also suggest a zonal variability that could be associated with solar tides
or vertically propagating orographic waves. Finally, the combination of our wind measurements from TNG/NICS,
IRTF/SpeX and Akatsuki images with previously published and based in data from 1978 to 2017 suggests variations
of up to 30 m s−1 in the winds at the lower clouds of the Venus nightside.
Keywords: planets and satellites: atmospheres, planets and satellites: terrestrial planets
1. INTRODUCTION
Corresponding author: Javier Peralta
[email protected]
The atmospheric circulation of Venus from the sur-
face up to the stratosphere is dominated by a retrograde
zonal superrotation (h.a. RZS) which, according to both
in situ (Counselman et al. 1980) and cloud tracking
measurements (S´anchez-Lavega et al. 2008), attains the
2
Peralta et al.
fastest wind speeds at the top of Venus's cloud layer
(Schubert 1983; Gierasch et al. 1997; S´anchez-Lavega
et al. 2017). The cloud layer is also where most of
the energy from the solar radiation is deposited (Lee
et al. 2015b), so a detailed characterization of the at-
mospheric circulation at multiple levels of the clouds
(S´anchez-Lavega et al. 2008; Peralta et al. 2017b) seems
essential to properly evaluate the sources and trans-
port of angular momentum in the atmosphere. The
wind speeds at the upper clouds of Venus (60 -- 70 km
above the surface) are customarily characterized with
cloud tracking on both day (Rossow et al. 1990; Belton
et al. 1991; Peralta et al. 2007; Limaye 2007; Kouyama
et al. 2013; Khatuntsev et al. 2013; Hueso et al. 2015;
Horinouchi et al. 2018) and night sides (Peralta et al.
2017c), while the winds at the deeper middle and lower
clouds (about 48 -- 60 km) have been evaluated following
cloud features visible on the dayside of the planet on
the albedo at near infrared wavelengths (Belton et al.
1991; Hueso et al. 2015; Khatuntsev et al. 2017) and in
the lower clouds on the nightside (S´anchez-Lavega et al.
2008; Hueso et al. 2012; Horinouchi et al. 2017b) thanks
to inhomogeneities in the opacity of the lower clouds
which can be observed at the infrared windows at 1.7
and 2.2-2.3 µm (Peralta et al. 2017a).
After the discovery of these infrared spectral windows
by Allen & Crawford (1984), and before the arrival
of Akatsuki (Nakamura et al. 2016), ESA's Venus Ex-
press mission (h.a. VEx) characterized the RZS at
the deeper clouds with unprecedented detail during
2006 -- 2008. However, constraints due to VEx's polar
orbit (Titov et al. 2006) and the long exposure times
required by the imaging spectrometer VIRTIS-M (Pic-
cioni et al. 2007) limited observations to the southern
hemisphere of Venus with lower quality results at equa-
torial and lower latitudes (S´anchez-Lavega et al. 2008;
Hueso et al. 2012; McGouldrick et al. 2012; Peralta et al.
2017b). As a result, the circulation at the level of the
deeper clouds is yet poorly characterized in the north-
ern hemisphere where only sparse measurements exist
from ground-based observations (Crisp et al. 1989, 1991;
Chanover et al. 1998; Limaye et al. 2006) and images
from the Near-Infrared Mapping Spectrometer (NIMS)
instrument during the flyby of NASA's Galileo in 1990
(Carlson et al. 1991). Since its orbit insertion in 2015
December, JAXA's Akatsuki orbiter has permitted an
invaluable opportunity to study this deeper atmospheric
circulation on both hemispheres thanks to its equatorial
orbit (Nakamura et al. 2016) and the images provided
by the 2-µm camera IR2 (Satoh et al. 2016).
In this work we present the first global measurements
of the wind speeds at the nocturnal lower clouds of
Venus during the first year of Akatsuki observations. A
description of the observations performed by the IR2
camera, the processing of the images, navigation correc-
tions and methods of cloud tracking are introduced in
section 2. The lower clouds' morphologies, the merid-
ional profiles of wind speeds and the possible relation
between morphologies and speeds are presented in sec-
tion 3. The wind speeds' dependence on the size and
opacity of the clouds, the local time and an explo-
ration of possible relations with the surface topography
is studied in section 4, while the temporal variation of
the winds of the lower clouds is presented in section 5.
Finally, the main conclusions of this work are presented
in section 6.
2. METHODS
After failing its originally-planned orbit insertion in
2010 December 7, Akatsuki orbited the Sun for 5 years
and was inserted with success into a Venus orbit in
2015 December (Nakamura et al. 2016). At present, the
orbiter performs a westward equatorial orbit with an
apoapsis of ∼360,000 km, periapsis ranging 1,000 -- 8,000
km, and a rotation period of about 10 days. Among
the payload, the camera IR2 was designed to sense the
deep clouds of Venus and infer information about the
atmospheric compounds below the cloud layer thanks
to its narrow-band filters centered at 1.74, 2.26 and
2.32 µm (Satoh et al. 2016). The scientific objectives
of these filters include the study of the morphology of
the clouds and their motions, the aerosols' properties or
the abundance of the CO below the clouds (Satoh et al.
2017). Because of the highly eccentric orbit of Akatsuki,
the spatial resolution of Venus in the IR2 images varies
from 74 -- 12 km per pixel for off-pericenter observations,
to 1.6 -- 0.2 km during pericentric ones (Nakamura et al.
2016). Unfortunately, the acquisition of IR2 images was
indefinitely interrupted in 2016 December 9 when the
electronic device controlling the IR1 and IR2 cameras
started to experience an unstable power consumption
that has persisted through the present time (Iwagami
et al. 2018).
In addition to the IR2 images, new wind measurements
have been obtained in this work with ground-based im-
ages of Venus acquired after the failure of the infrared
channel of the VIRTIS-M (Hueso et al. 2012) instrument
on VEx and before and after the time period covered by
Akatsuki IR2 observations. The Near Infrared Camera
Spectrometer (NICS) (Baffa et al. 2001) at the Ital-
Nightside Winds of Venus with Akatsuki/IR2
3
ian National Telescope Galileo (TNG) at La Palma
(Canary Islands, Spain) was used to acquire images of
the nightside of Venus during 2012 July 11 -- 13 (Mota
Machado et al. 2016), while the Medium-Resolution 0.8 --
5.5 Micron Spectrograph and Imager (SpeX) (Rayner
et al. 2003) at the 3-m National Aeronautics and Space
Administration Infrared Telescope Facility (IRTF) was
used to provide images of Venus with the K-continuum
filter in 2015 September and 2017 January-February
(Lee et al. 2017a). The wind measurements from these
ground-based images will be presented in section 5.
2.1. Image processing
IR2 acquired a total of 1,671 images1 of the night-
side of Venus with the filters 1.74, 2.26 and 2.32 µm
(Satoh et al. 2017) among which ∼1,370 were regarded
as suitable for cloud tracking. Satoh et al. (2017, fig. 8a
therein) reported a problem of light contamination in
the IR2 images, consisting of the presence of halation
rings and a cross pattern extending both horizontally
and vertically around the saturated dayside of Venus
and spreading with multiple reflections along the PtSi
detector. Since this contamination is more reduced in
the images taken with the 2.26-µm filter, we restricted
our cloud tracking study to this data set composed of
466 images (although additional sets with 1.74-µm im-
ages were used in the case of automated cloud tracking).
The calibration version of the IR2 images used in this
work ("v20170601") does not include any of the cor-
rections for the light contamination proposed by Satoh
et al. (2017) and an alternative image processing tech-
nique was applied. This consisted in an adjustment
of the brightness/contrast, followed by sharpening the
images with an unsharp-mask technique, and finishing
with the application of adaptive histogram equalization
(see examples in Fig. 1). Images acquired with ground-
based instruments at TNG and IRTF also suffered from
contamination from the illuminated side of the planet
and were processed similarly. Some examples of these
images are shown in section 5.
2.2. Navigation of IR2 and ground-based images
Uncertainties such as the thermal distortion affecting
the Akatsuki spacecraft and the cameras onboard pre-
vent high accuracy in the navigation of the Venus images
at present, so additional corrections in the navigation
are yet required (Ogohara et al. 2017; Satoh et al. 2017).
In recently published studies with Akatsuki (Fukuhara
et al. 2017; Horinouchi et al. 2017b; Lee et al. 2017b;
Horinouchi et al. 2018) the navigation of the images
was corrected with an algorithm able to perform an
ellipse fitting from an automatic determination of the
planetary limb pixels (Ogohara et al. 2012, 2017). This
automatized identification of the limb becomes rather
uncertain in many of the IR2 nightside images because
of the light contamination previously described, and
also due to the frequent darkening of the planetary limb
because of the strong variability of the clouds' opacity
at lower latitudes McGouldrick et al. (2008); Satoh et al.
(2017). Instead, we coded an interactive tool inspired
on the software WinJupos (Hahn & Jacquesson 2012)
which allows the interactive adjustment of the position,
size and orientation of the planet's grid2, using as refer-
ence four locations of the limb chosen by the user (see
Fig. 2). The visualization of the limb was improved by
interactively modifying the brightness and contrast in
each image. In the case of the IR2 images, the position
of the grid was adjusted with a precision of 1/10th of
pixel, its size was increased in some cases (less than
1.3% in all the cases), while the orientation of the grid
required no corrections. Regarding the ground-based
images from NICS and SpeX, these were navigated us-
ing NASA's SPICE kernels (Acton 1996; Folkner et al.
2009), both the position and orientation of the grid were
adjusted, and the predicted size was accurate enough to
require no further corrections.
2.3. Techniques for Cloud Tracking
After correcting the navigation and processing the im-
ages, these were geometrically projected onto equirect-
angular (cylindrical) geometry with an angular resolu-
tion similar to the best resolution in the original images.
In those cases of nightside images acquired when Akat-
suki was closer to its pericenter, the better spatial reso-
lution enabled measuring wind speeds at polar latitudes
and azimuthal equidistant (polar) projections were per-
formed too using an angular resolution similar to that
found at latitudes of about 70◦ in the original image.
Figure 3 shows examples of the original observations,
their navigation after corrections and cylindrical and
polar maps. For the measurement of wind speeds, two
different techniques of cloud tracking were employed:
(a) a manual method applied to the full data set of IR2
and ground-based images (see Table 1), consisting of a
manual search of the cloud tracer followed by a fine ad-
1 Available at: http://darts.isas.jaxa.jp/planet/project/
2 The corrected geometry files are available from the corre-
akatsuki/
sponding author on reasonable request.
4
Peralta et al.
Figure 1. Examples of IR2/2.26-µm images before and after the processing procedure described in this work: (A) IR2 image
acquired in March 25, 11:33 UT, and (B) IR2 image acquired in August 25, 04:03 UT. Both images have been oriented so that
the north-south-west-east of Venus appears approximately oriented up-down-left-right in the image.
Figure 2. Example of the correction on the navigation of Akatsuki images using as reference a set of locations of the planetary
limb selected by the user. The original navigation grid and the corrected one are displayed on the left and right sides, respectively.
justment using an automatic template matching which
is visually accepted or rejected by the human operator
(similarly as performed by Hueso et al. 2015); and (b)
a fully-automatic method used for images taken in 2016
July and August (see Table 1) and which applies the
relaxation labeling technique (Ikegawa & Horinouchi
2016; Horinouchi et al. 2017a).
Wind measurements are obtained by comparing the
position on maps of cloud features that can be identified
in two consecutive images with a given time difference.
We obtained wind measurements with a manual method
in which we perform an automatic template matching
with a phase-correlation technique that has been also
applied for cloud tracking on the Earth (Leese et al.
1970; Jun & Fengxian 1992; Humblot et al. 2005; Huang
et al. 2012). We call this technique manual because a
Nightside Winds of Venus with Akatsuki/IR2
5
Figure 3. Examples of images acquired with the filter 2.26-µm by Akatsuki/IR2 during the year 2016 (A -- G) and ground-based
observations at 2.32-µm from TNG/NICS and IRTF/SpeX (H -- J): (A) 2016 October 10, 09:23 UT, (B) 2016 August 13, 00:33
UT, (C) 2016 July 11, 22:03 UT, (D) 2016 July 22, 14:33 UT, (E) 2016 October 19, 14:33 UT, (F) cylindrical and (G) polar
projections for 2016 March 25, 07:33 UT (original shown in Fig. 6A), (H) 2012 July 11, 05:19 UT by TNG/NICS, (I) 2015
September 28, 16:03 UT by IRTF/SpeX, and (J) 2017 February 10, 02:31 UT by IRTF/SpeX. In the case of the polar projection
(G), the northern and southern hemispheres are displayed above and below, respectively, with the geographical north/south
pole located in the center.
6
Peralta et al.
human user selects a region in the first image to obtain
its best match in the second image and validates (or not)
the final result. The phase-correlation permits to obtain
the translation between two images shifted relative to
each other (Kuglin 1975; Samritjiarapon & Chitsobhuk
2008), relying on a frequency-domain representation of
the images calculated with the Fast Fourier Transform
that enables to infer this shift from the location of a
peak in a cross-correlogram. Figure 4 shows an example
of the use of this technique where the narrow peak in
the phase cross-correlogram is shown for a well identified
tracer. A more complete explanation about the phase-
correlation is provided by Kuglin (1976). Compared
to the standard correlation, its performance is faster,
and it is less sensitive to random noise and illumina-
tion conditions (Ahmed & Jafri 2008). Even though
the phase-correlation can potentially register sub-pixel
displacements (Reddy & Chatterji 1996; Foroosh et al.
2002), we did not test this capability in this work and
only accounted for displacements expressed as integers
of pixels. Besides, since the boundaries of any image
imply discontinuities in the signal that can introduce a
noisier result, a zero-padding is advisable before apply-
ing the FFT (Ahmed & Jafri 2008). For this reason,
the initial and final templates (ranging sizes of 32×32
to 48×48 pixels) were convolved with a Hanning win-
dow with a width of 0.7 before the FFT was applied.
When no clear peak could be found with the phase cor-
relation, or when the tracer identification was judged
as not satisfactory for the human operator running the
analysis, the wind measurement was rejected and the
template matching was performed manually. Each of
these manual measurements was verified by a human
operator looking at visual reports as the one shown in
Figure 4.
A fully automatic technique based on classical image
correlation for identifying the cloud tracers motions was
also used by Horinouchi et al. (2017a,b) and we will
show here detail report of these measurements. This
technique is similar to those used in many previous
studies of Venus cloud dynamics (Rossow et al. 1990;
Kouyama et al. 2012; Khatuntsev et al. 2017; Hori-
nouchi et al. 2018), but in this case the cross-correlation
between images is computed along a sequence of images
to estimate the horizontal velocity at a specific location
(Horinouchi et al. 2017b, see Methods therein). This
method was applied to images of the lower clouds ac-
quired in 2.26 and 1.74-µm. All the IR2 images were
projected onto equirectangular geometry with a fixed
angular resolution of 0.125◦ per pixel regardless of the
resolution of the original images (Ogohara et al. 2017).
Figure 4. Example of positive identification of a cloud
tracer in a pair of IR2 images using phase-correlation. Note
that when a match is successful (see templates), a single and
clear peak -- ideally Kronecker delta -- is apparent in the cross-
correlogram located in the corner right-up. The inferred
wind speeds are compared with a reference wind profile from
VEx/VIRTIS-M (Hueso et al. 2012) in the two graphs below,
while detailed information is provided on the right side below
the correlogram. A complete report of clouds' matches with
phase correlation can be found in the Supplemental Material
(see Appendix A).
Image processing was applied after the projection and
consisted on a two-dimensional bandpass filter with
Gaussian functions with the sigma values 0.25◦ for low
pass and 3◦ for high pass with latitude and longitude.
Finally, the template size for cloud tracking was set to
a fixed value of 60×60 pixels for all the images.
3. MERIDIONAL PROFILES OF WIND SPEEDS
A total of 2,947 wind vectors were obtained with the
manual method for the nightside images of the IR2
camera for the full period covering during 2016 from
March 22 (when pairs of IR2 images were taken within
the same day for the first time) until October 31. The
spatial resolution of the images ranged 10 -- 80 km per
pixel and we selected pairs of images with time sepa-
rations ranging from 1 hour up to 22 hours (in special
cases of clouds displaying very small deformations over
large time scales). The size of the cloud tracers com-
prise dimensions ranging 510 -- 2,550 km, depending on
the spatial resolution selected for the geometrical pro-
jections.
Images acquired over 2016 April, July and
Nightside Winds of Venus with Akatsuki/IR2
7
August were also independently measured with the
fully-automatic method described above resulting in
149,033 wind measurements from the tracking of cloud
features with a fixed size of 790 km. The errors for
the wind speeds obtained with manual tracking were
calculated from the spatial resolution and the time in-
terval between the images as explicated by Bevington &
Robinson (1992), while in the case of automatic cloud
tracking the calculation of the errors are explained by
Ikegawa & Horinouchi (2016).
Satoh et al. (2017, fig. 11b therein) and Limaye et al.
(2018, fig. 15 therein) show Akatsuki IR2 images of
strong cloud discontinuities in the lower clouds of Venus.
These discontinuities propagate faster than the back-
ground zonal flow and were suspected to be the mani-
festation of waves rather than passive tracers. For this
reason, we discarded the motions of these strong opac-
ity discontinuities from our study of the global wind
motions. In the case of the automatic cloud tracking,
we ruled out all the measurements obtained from im-
age pairs where this equatorial cloud discontinuity was
apparent. As a result of this filtering, the number of
wind vectors with the manual and automatic methods
was reduced to 2,277 and 101,882 respectively. Since
the automatic method does not provide a registry of the
cloud morphologies tracked, this filtering also removed
all automatic-generated wind vectors during April 2016.
Table 1 summarizes the results obtained with both tech-
niques in different time periods and their coverage over
the planet. The full report of wind measurements from
the 2.26-µm IR2 images without filtering out specific
features and obtained with the manual method can be
found in the Supplemental Material (see Appendix A),
including a table with the complete data set of wind
measurements with manual tracking, geometrical pro-
jections, animations and detailed template matching
results.
3.1. Morphologies of the nightside clouds
Panels A -- G in Figure 3 exhibits a sample of the
clouds' morphologies observed in the IR2/2.26-µm im-
ages. Pioneering ground-based observations at these
wavelengths (Allen & Crawford 1984; Crisp et al. 1991)
show similar global characteristics as Akatsuki IR2 im-
ages. The deeper clouds of Venus nightside are normally
characterized by a dark band with high opacity clouds
at low latitudes, while the mid-latitudes are dominated
by brighter bands with lower opacity (see Figs. 3A -- B).
Prior to Akatsuki, the mid-latitude bands appeared as
almost featureless (Crisp et al. 1991; Limaye et al. 2006;
Hueso et al. 2012). However, the higher spatial resolu-
tion of the Akatsuki IR2 images reveals on them subtle
though distinguishable wisps and patches (Figs. 3F -- G)
sometimes invaded by broad bands of clouds with higher
opacity (Fig. 3B), or unusual sharp dark spirals tilted
relative to the latitude parallels (Figs. 3C -- D) which
spread thousands of kilometres from latitudes higher
than 30◦ towards equatorial ones (Horinouchi et al.
2017b; Satoh et al. 2017; Limaye et al. 2018).
In the IR2 images, the clouds' opacity at lower lati-
tudes exhibits higher variability than at mid-latitudes
(Satoh et al. 2017; Limaye et al. 2018), confirming pre-
vious findings with ground-based observations (Crisp
et al. 1991; Limaye et al. 2006; Tavenner et al. 2008;
Mota Machado et al. 2016) and during the VEx mis-
sion (Hueso et al. 2012; McGouldrick et al. 2012; Mc-
Gouldrick & Tsang 2017). Dark areas of high opacity
normally dominate lower latitudes, while their bound-
ary with the mid-latitude bright bands exhibits complex
cloud features (Horinouchi et al. 2017b; Limaye et al.
2018), like small vortices or abundant bright swirls at
around 30◦ which sometimes adopt shapes suggestive
of shear instabilities (Fig. 3B,C,E), similar to those
found at the nightside upper clouds at ∼65 km with
3.8-µm VEx images(Peralta et al. 2017c). The polar
projection in Fig. 3G clearly exhibits an example of the
frequent episodes of hemispherical asymmetry for the
Venus clouds' morphology. A more complete survey of
the clouds' morphologies apparent in the images by the
Akatsuki/IR2 camera will be presented elsewhere.
3.2. Latitudinal profiles of the winds and relation with
clouds
Figure 5 shows zonally-averaged profiles of
the
manually-tracked winds at the lower clouds during the
first year of Akatsuki observations. Bins of 5◦ of lati-
tude were considered to calculate average wind speeds
from March to October 2016. The profiles of IR2 are
compared with VEx/VIRTIS-M results from 2006 April
to 2008 August (Hueso et al. 2012). With regards
to the zonal component of the wind (Fig. 5A), the
mean profile obtained with IR2 is faster than VIRTIS-
M measurements by 10 ms−1. The zonal winds from
IR2 observations show symmetric profiles between both
hemispheres and, despite the higher dispersion at higher
latitudes due to the poorer spatial resolution of the IR2
images close to the poles, they show for the first time
the decay of the winds towards the poles on both hemi-
spheres simultaneously. This decay starts at about 60◦
and, compared to VEx results, it seems more abrupt,
8
Peralta et al.
Table 1. Coverage of Akatsuki/IR2 Winds during year 2016
Month
March
April
June
July
Dates*
22,23,25 -- 30
15
20
1,11,12,22
August
2,9,10,13,15 -- 17,
18,19,20,21,22,25,
26,27,28,29,30
Latitudes
61◦N -- 60◦S
43◦N -- 63◦S
40◦N -- 21◦N
72◦N -- 69◦S
58◦N -- 58◦S
Local Times Longitudes Wind Vectors**
23h -- 05h
01h -- 06h
18h -- 20h
18h -- 01h
18h -- 03h
225◦ -- 308◦
281◦ -- 348◦
264◦ -- 289◦
250◦ -- 23◦
322◦ -- 129◦
249
35+306
12
308+6920
1057+94656
September
4 -- 6,15,26,27
October
2 -- 7,10 -- 17,
59◦N -- 60◦S
56◦N -- 72◦S
19h -- 05h
20h -- 05h
28◦ -- 188◦
114◦ -- 270◦
347
939
19 -- 27,31
∗Bold characters stand for dates with wind measurements from both manual and fully-automatic
∗∗Normal and bold characters stand for number of wind measurements obtained with the manual
methods. Normal characters are used for dates with only manual measurements.
and fully-automatic methods, respectively.
while it keeps similarities with wind profiles obtained
in 2004 with ground-ground observations (Limaye et al.
2006, fig. 5 therein), albeit, with higher errors. However,
we cannot rule out that these differences at subpolar lat-
itudes with respect to VEx/VIRTIS-M subpolar winds
are not caused by the worse spatial resolution at polar
latitudes and the low number of measurements in the
Akatsuki IR2 images (Fig. 5C). Regarding the merid-
ional component of the wind (Fig. 5B), the results of
2016 and 2006 -- 2008 are in good agreement, confirming
the absence of clear global motions in the meridional
circulation of the nocturnal lower clouds.
Wind variability is also apparent during 2016 (see
Fig. 6). The profile obtained in 2016 March 25 (Fig. 6A)
corresponds to winds that are approximately constant
between the equator and mid-latitudes with small
meridional shear. This seems to be most frequent case
found at the lower clouds during the Akatsuki obser-
vations and also during the VEx mission (Hueso et al.
2012, fig. 7 therein). Zonal wind profiles on other dates
like the July 11 (Figs. 6B) and October 13 (6C) exhibit
more intense equatorial zonal speeds, first-time reported
during the Galileo flyby (Crisp et al. 1991, fig. 4 therein)
and later identified as recurrent episodes of jets at the
equator (Horinouchi et al. 2017b). There is a reasonable
correspondence between this local intensification of the
zonal speeds and the presence of features resembling
shear instabilities (Fig. 6B) or sharp opacity disconti-
nuities at the equator (Fig. 6C). The high dispersion in
the horizontal speeds also suggests that these jets may
be apparent only in a rather longitudinally narrow area.
The profile during July 11 (Fig. 6B) shows that these
jets can sometimes show up at northern latitudes, as
also reported in 1990 during the Galileo flyby (Carlson
et al. 1991, fig. 6 therein) and in 1996 from observations
at the Apache Point Observatory (Chanover et al. 1998,
fig. 7 therein).
4. SOURCES OF VARIABILITY FOR THE WIND
SPEEDS AT THE LOWER CLOUDS.
In this section, we explore whether the size of the fea-
tures tracked, the opacity of the clouds, the uncertain-
ties in their vertical sensing, or a dependence on the lo-
cal time and surface elevations, can help to explain the
variability observed in the winds of the nightside lower
clouds.
4.1. Effect of opacity and size of the clouds
It is yet unclear whether the episodes of faster zonal
speeds shown in Fig. 6 may be related to horizontal
and/or vertical gradients of the zonal wind, or if the
bright and dark clouds observed on the nightside at
1.74, 2.26 and 2.32 µm are moving at the same or
at slightly different vertical levels. Early interpreta-
tions of Venus IR features considered that these clouds'
contrasts might be caused either by scattered sunlight
leaking in from the dayside along altitudes with low
absorption in the CO2 windows, or that dark areas
Nightside Winds of Venus with Akatsuki/IR2
9
Figure 5. Mean profiles of the winds at the nocturnal middle-to-lower clouds of Venus as measured from March to October
2016 with the manual method using the Akatsuki/IR2 2.26-µm images (dark red line). These profiles are compared with the
winds measured with 1.74-µm images acquired with VEx/VIRTIS-M (blue line) between 2006 April and 2008 August (Hueso
et al. 2012). The profiles for the zonal and zonal and meridional components are shown in panels (A) and (B), respectively.
Average values shown here were calculated for bins of latitude of 5◦, and the number of measurements used in each bin are
displayed in panel C.The shadowed areas stand for the standard deviation at every latitude bin when this is larger than the
measurement error computed from the image resolution and time differences in the image pair.
corresponded to opaque clouds lying in a broken layer
at certain altitude above the brighter regions and ra-
diate as blackbodies at the lower temperature of their
tops (Allen & Crawford 1984). Allen (1987) discarded
both interpretations, confirming that the existence of
important differences in the vertical elevation of bright
and dark clouds was inconsistent with the small varia-
tions that dark/bright areas exhibit in CO absorptions
affecting the wide 2.2-2.3 µm band. Therefore, the
dark and bright regions should correspond to horizontal
inhomogeneities in the clouds' opacity to the deeper
background thermal emission (Allen 1987; Crisp et al.
1989), although certain discrepancies of altitude be-
tween dark and bright clouds cannot be discarded since
these inhomogeneities can be caused by variations in the
size and distribution of the particles within the lower
and middle clouds (McGouldrick & Toon 2008).
Assuming a range of 50 -- 60 km for the altitudes sensed
at the relevant wavelengths (McGouldrick et al. 2008),
a comparison with in situ wind profiles from descending
probes (Gierasch et al. 1997; Counselman et al. 1980;
Moroz & Zasova 1997) indicates that variations of up
to 30 m s−1 could be explained in terms of the vertical
shear (Peralta et al. 2017c, fig. 3 therein), what is also
consistent with the magnitude of the reported jets (Hori-
nouchi et al. 2017b, fig. 2b therein). Crisp et al. (1991)
obtained a discrepancy between velocities of large dark
clouds and the smaller markings (sizes ranging 400 -- 1000
km), and also suggested that these might be produced
at different altitudes. Figure 7 displays the distribu-
tion of zonal speeds obtained with the manual method
compared with the averaged radiance (Fig. 7B) and the
size of the cloud tracers between 50◦N -- 50◦S (Fig. 7A) in
the IR2 2.26-µm images. No dependence is apparent be-
tween the zonal velocities and these parameters. Since
the low latitudes frequently exhibit cloud patterns that
resemble shear instabilities for a wide range of scales
(see Figs. 3C -- E; Limaye et al. 2018, fig. 10 therein),
the generation of jets due to meridional gradients in the
horizontal winds seems probable. McGouldrick & Toon
(2007, 2008) showed that large-scale dynamics can also
10
Peralta et al.
Figure 6. Variability in clouds' morphology (left column)
and profiles of zonal and meridional winds (center and right
column) during the Akatsuki mission in 2016: (A) winds
during March 25, exhibiting the standard profile of constant
zonal speeds; (B) winds during July 11, displaying a zonal
jet at ∼20◦N; and (C) winds during October 13, exhibiting a
strong jet at the equator and higher dispersion in the south-
ern hemisphere. Reference profiles built from winds during
the VEx mission (Hueso et al. 2012) and plot in a symmetric
way for the two hemispheres are displayed in cyan.
explain the strong variations in the cloud opacity, and
that even weak downwelling is able to produce optical-
depth holes in the clouds.
4.2. Local time dependence.
Since most of the sunlight absorption occurs on Venus
within the clouds' layer, solar tides are expected to be
excited and carry momentum away upwards and down-
wards from the region of excitation, thus accelerating
the atmosphere westwards and contributing to the RZS
(Gierasch et al. 1997; S´anchez-Lavega et al. 2017). The
effect of the solar tides have been unambiguously de-
tected on both zonal and meridional components of
the winds at the upper clouds at 65 -- 70 km (Rossow
Figure 7. Histograms and scatter plot of the wind measure-
ments acquired with the manual method and Akatsuki/IR2
images. The panel (A) shows histograms of the zonal ve-
locities for two ranges of spatial scales of cloud tracers. The
panel (B) exhibits the values of zonal wind speeds in terms
of the mean radiance (red dots). The radiance is given for
the calibration version "v20170601" of the IR2 2.26-µm im-
ages, and corresponds to the radiance averaged inside the
template.
et al. 1990; Limaye 2007; S´anchez-Lavega et al. 2008;
Kouyama et al. 2012; Peralta et al. 2012; Khatuntsev
et al. 2013; Hueso et al. 2015), but no evident influence
has been found for the middle-to-lower clouds (50 --
60 km) for either day (Hueso et al. 2015; Khatuntsev
et al. 2017) or night (Hueso et al. 2012, fig. 6 therein).
Khatuntsev et al. (2017) argued that this negligible ef-
fect of the solar tides could be related to the absence of
the unknown absorber downward of the middle clouds
(which is responsible for most of the solar heat deposi-
tion on Venus).
Figure 8 allows to study the local time dependence for
the winds at the nightside lower clouds of Venus during
2016. In agreement with VEx results during 2006 -- 2008,
no local time dependence is apparent in the meridional
Nightside Winds of Venus with Akatsuki/IR2
11
winds (Fig. 8B). However, the zonal component of the
wind speed displays a local increase of the zonal speeds
between 19 -- 22 LT, followed by a gradual decrease to-
wards late nightside local times (decrease of ∼10 m
s−1 between 19 LT and midnight). Hueso et al. (2012)
also reported hints of faster retrograde winds close to
dawn during 2006 -- 2008, although this local time effect
was weaker and disregarded due to the small number of
measurements in this area. Recent results from Venus
General Circulation Models (h.a. GCMs) predict that
the diurnal tide should be also apparent on the zonal
wind down to the middle clouds at ∼60 km with the
slowest speeds centered at the evening terminator (Tak-
agi et al. 2018, fig. 3a therein). Better agreement is
found for the GCM results at the lower clouds (∼50
km), where the influence of the diurnal tide is predicted
to weaken but a local maximum of about 10 m s−1 is
also found in the early night and before the midnight
(Takagi et al. 2018, fig. 4a therein).
4.3. Longitudinal dependence.
Evidences on how the surface topography may be in-
fluencing the atmospheric circulation through the exci-
tation of atmospheric stationary waves (lee waves) have
been accumulating. Vertical disturbances experienced
by the VEGA-2 balloon over Aphrodite terra (Blamont
et al. 1986), strong asymmetries of the water vapor over
certain geographical locations (Fedorova et al. 2016), or
the finding of multiple stationary waves at the upper
clouds of Venus which are strongly correlated with the
surface elevations (Peralta et al. 2017c; Fukuhara et al.
2017; Kouyama et al. 2017) strongly suggest surface
effects on the upper atmosphere. Results of cloud track-
ing in VEx/VMC dayside images seem to support that
the wind speeds at the cloud tops are decelerated as
they pass over Aphrodite terra and Atla regio (Bertaux
et al. 2016, figs. 4 and 6 therein). Other works have
reported indications of an effect of surface elevations for
the winds at the middle cloud (Khatuntsev et al. 2017,
fig. 14 therein) and over the oxygen airglow patterns
(Gorinov et al. 2018).
Peralta et al. (2017c) discovered that during 2006 --
2008 the RZS at the nocturnal upper clouds exhibited
a higher variability compared to the dayside, but the
geographical coverage of the wind measurements was
insufficient to confirm a correlated effect with the sur-
face elevations. In the case of the lower clouds on the
nightside, stationary waves are paradoxically missing
in the clouds' opacity and no longitudinal dependence
had been reported for the wind speeds (a reanalysis
of VIRTIS-M winds for the lower clouds resulted in-
conclusive due to the lack of enough data to separate
local time effects from time variability and the elusive
surface dependence). Figure 9 displays the dependence
on the longitude and latitude for the manual wind mea-
surements with the Akatsuki IR2 images. While no
clear effect is found for the meridional component of
the wind (Fig. 9B), the westward windspeeds exhibit a
local maximum at low latitudes over longitudes ranging
300◦ -- 120◦ (fig. 9A).
Unfortunately, separating the effects of the longitudi-
nal and local time dependences is not feasible with our
results, since the observed velocity disturbances are just
marginally larger than the corresponding errors and the
incomplete coverage of our wind measurements. Fig-
ure 10 shows the distribution of the mean zonal speeds
(panel 10A) and the number of measurements (10B) in
terms of local time and longitude. To avoid the expected
decrease of the zonal winds towards the poles, only the
Akatsuki/IR2 wind measurements between 50◦N -- 50◦S
were considered. Panel 10B shows that the number of
wind measurements during the year 2016 is irregularly
distributed for both longitude and local time param-
eters, preventing a confirmation of either a local time
dependence and/or influence of surface elevations.
However, the influence of the surface elevations over
the zonal winds at the nightside lower clouds might be
yet regarded as controversial. Published results for the
zonal winds on the dayside were originated from a huge
set of wind measurements with noticeable dispersion
(Bertaux et al. 2016, fig. 2 therein), and authors do not
report filtering out other important sources of variabil-
ity such as transient waves or the local time dependence.
In the case of zonal wind speeds obtained with Akat-
suki/UVI images, Horinouchi et al. (2018) characterized
and removed the local time dependence on the zonal
winds at the dayside upper clouds. As a result, no effect
of surface elevations was found for the dayside winds
(Horinouchi et al. 2018, fig. 14b therein).
In the case
of our results for the nightside lower clouds, the local
maximum extends mostly over rather plain areas and
do not seem well correlated with the surface elevations
at lower latitudes. Moreover, our results seem incon-
sistent with those on the dayside reported by Bertaux
et al. (2016) and Khatuntsev et al. (2017) since, con-
versely to the zonal wind speeds on the dayside from
VEx/VMC which exhibit a local minimum with slower
speeds extending along 0◦ -- 200◦ (Bertaux et al. 2016,
fig. 3 therein), our westward winds on the nightside dis-
12
Peralta et al.
Figure 8. Local time dependence for the zonal winds obtained with the manual method on the Akatsuki/IR2-2.26-µm images.
Bins of 10◦ in latitude and 40 minutes in local time were used. Panels (A) and (B) display the values of zonal and meridional
components of the wind averaged in each bin; panel (C) shows the distribution of wind measurements with dots representing
individual measurements; (D) and (E) display the standard deviation of the measurements in each bin; and (F) shows the
meridional profiles of zonal winds at different intervals of local time.
play a local maximum between longitudes 300◦ -- 120◦.
4.4. Results from automatically retrieved winds during
2016 July and August
We now explore possible winds dependencies with the
factors shown above for our wind results obtained in
the automatic cloud correlation technique and firstly
reported by Horinouchi et al. (2017b). The dependence
with both local time and longitude for the winds ob-
tained with the fully-automatic method are displayed in
Figure 11. Since for these wind speeds only IR2 images
obtained during 2016 July and August were used (see
Table 1), the coverage is more limited than in the case of
the manual method, and the dispersion is smaller than
for the manual results (see figs. 8D,8E,9D and 9E).
Except for the case of the meridional winds where
no clear dependence with the local time and longitude
is found, the zonal winds with automatic correlation
are not fully consistent with that from manual wind
measurements, and no influence over the zonal speeds is
observed at any range of local time or longitude. Even
though some discrepancies were expected between both
cloud tracking methods and the different techniques to
correct the navigation of the IR2 images, the homo-
geneity found for the zonal winds at low latitudes might
be explained due to the recurrence of the equatorial jet
during August 2016 (Horinouchi et al. 2017b), which
would mask a weaker dependence on local time and/or
longitude.
4.5. Combined results of Akatsuki and Venus Express
Nightside Winds of Venus with Akatsuki/IR2
13
Figure 9. Local time dependence for the zonal winds obtained with the manual method on the Akatsuki/IR2-2.26-µm images.
Bins of 10◦ and 40' were considered for the latitude and local time, respectively. Panels (A) and (B) display the values of zonal
and meridional components of the wind averaged in each bin; panel (C) shows the distribution of wind measurements; (D) and
(E) display the error (standard deviation) in each bin; and (F) exhibits the surface elevation of Venus as a comparison.
We also combined the manual wind speeds at the
nocturnal lower clouds with Akatsuki/IR2 in the year
2016 with those obtained by Hueso et al. (2012) with
VEx/VIRTIS-M from 2006 to 2008. Since the mean
zonal winds during 2016 were faster than during the
VEx mission (see Fig. 5A), the difference between
the zonal averages of both datasets equatorward of
midlatitudes was suppressed by multiplying by a cor-
recting factor of 1.06 the zonal speeds obtained with
the VEx/VIRTIS-M images, while those from Akat-
suki/IR2 images were divided by 1.06. As a result, a
correction of ∼4 m s−1 was introduced in both sets
of winds. Also, these two datasets have a comparable
number of points and both are based on cloud tracking
inspected/validated by human operators. The result of
their combination is presented in the figure 12. The
dependence with the local time (panels 12A -- B, E -- F)
confirms the equatorial maximum of the zonal speeds
between local hours 18h -- 22h, while this dependence
clearly decays towards higher latitudes. On the other
hand, the meridional component of the wind displays no
apparent dependence at low latitudes, while at higher
latitudes of the southern hemisphere (50◦S -- 90◦S) and
between 21h -- 22h it exhibits an equatorward accelera-
tion (panel 12B). These results, again, suggest that the
solar tide detected on the winds of the upper clouds
(Limaye 2007; Peralta et al. 2012) might be able to
propagate downwards to the middle clouds as predicted
by some Venus GCMs (Takagi et al. 2018). Regarding
the dependence with longitude and surface elevations,
the local maximum between longitudes 300◦ -- 120◦ re-
ported in subsection 4.3 for the westward windspeeds is
observed (panel 12C), while the meridional component
exhibits no clear influence from the surface (panel 12D).
14
Peralta et al.
the clouds' dynamics might experience cyclic changes
with a time scale of 5 -- 10 years. This conclusion was
based on the periods of presence/absence of the 4 and
5-day wave modes, while during the VEx mission Lee
et al. (2015a, fig. 18 therein) reported unexplained pe-
riods of 154,275,357 and 560 days on the 365-nm albedo.
Long-term trends were also observed for the dayside
winds of the upper clouds during the VEx mission. Be-
tween 2006 -- 2013 the mean zonal velocity measured with
ultraviolet images from the VMC camera displayed an
increase from 80 to 100 m s−1 at about 20◦S (Khatunt-
sev et al. 2013, fig. 14 therein), while a similar increase
trend was independently confirmed in this period with
cloud tracking in VIRTIS-M images (Hueso et al. 2015).
Kouyama et al. (2013) analyzed VEx/VMC images find-
ing that the zonal winds equatorward of 30◦S exhibited
a periodical perturbation of ∼10 m s−1 every 257±2
terrestrial days which was ultimately interpreted as
centrifugal waves (Peralta et al. 2014, fig. 4 therein).
Measurements with Akatsuki/UVI images have also
confirmed zonal windspeeds variations with time scales
of ∼100 days between December 2015 and March 2017
(Horinouchi et al. 2018, figs. 8 -- 10 therein). Numerical
simulations with GCMs have also been used to predict
decadal variation in the zonal winds in Venus (Parish
et al. 2011).
The long-term variability of the winds at the level of the
lower clouds was studied during the VEx mission on the
nightside with VIRTIS-M by Hueso et al. (2012) and
on the dayside with VMC by Khatuntsev et al. (2017).
During the years 2006 -- 2008, the nightside winds at the
deeper clouds showed at subpolar latitudes stronger
variability than at lower ones (Hueso et al. 2012, figs. 7 --
8 therein), while a comparison between the zonal winds
during the Galileo flyby in 1990 and the 2-year average
from VEx do not indicate noticeable variations except
for the northern jet reported from the Galileo/NIMS
images (S´anchez-Lavega et al. 2017, fig. 3 therein).
McGouldrick & Tsang (2017) reported an oscillation of
approximately 150 days apparent between 30◦S -- 60◦S on
the nightside clouds' radiance of the 1.74-µm VIRTIS-
M images, with this time scale being consistent with
the cycle of cloud formation and evolution driven by
the radiative dynamical feedback and gravitational set-
tling of clouds' particles. The analysis of the dayside
winds with VMC from 2006 December to 2013 August
(∼1,200 days) revealed long-term variations on both
components of the wind at ∼20◦S, with the meridional
winds exhibiting a gradual increase until doubling its
magnitude, while the zonal winds seemed subject to
Figure 10. Distribution as a function of Longitude and Lo-
cal Time of the wind measurements acquired with the man-
ual method and Akatsuki/IR2 images between 50◦N -- 50◦S.
The panel (A) displays the distribution of the zonal wind
speeds. Panel (B) shows the distribution of the number of
measurements.
5. TIME EVOLUTION OF THE WINDS
Results from past space missions to Venus and com-
parison between them have provided evidence of long-
term variability affecting the atmospheric dynamics
at the upper clouds of Venus (S´anchez-Lavega et al.
2017, fig. 2 therein). Rossow et al. (1990) reported
that during the Pioneer Venus zonal winds at equa-
tor and mid-latitudes exhibited variations ranging 5 -- 8
m s−1 over time spanning 1 -- 6 years, while the inten-
sity of the poleward circulation seemed also subject to
variability. From the analysis of the ultraviolet albedo
of the cloud tops in different missions, del Genio &
Rossow (1990) suggested that between 1979 and 1986
Nightside Winds of Venus with Akatsuki/IR2
15
Figure 11. Contour maps for the dependence of zonal and meridional winds obtained with the full-automatic method in the
Akatsuki/IR2 images. Averages (panels A,C,E,G) and standard deviation (B,D,F,H).The larger number of wind measurements
allows to calculate these maps with bins of latitude, longitude and local time of 3.3◦, 3.3◦ and 13 minutes respectively, or three
times smaller than in the manual tracking results. However, the spatial coverage over local times and longitudes is more limited
in this case due to the data coming from a smaller number of dates.
an apparent oscillation of ∼3 years (Khatuntsev et al.
2017, fig. 11 therein).
In order to perform an analysis of the variability of
the zonal winds at the nightside lower clouds using an
even longer time scale, we decided to combine data not
only from Akatsuki/IR2 and VEx/VIRTIS, but also
from ground-based observations and in situ measure-
ments. Figure 13 shows the variability of zonal winds
at low latitudes (30◦S -- 30◦N) of the nightside lower
clouds from 1978 December to 2017 February (span-
ning 38 years). The zonal winds during 1978 December
(Fig. 13A) and 1985 June (Fig. 13D) correspond to the
averages of instantaneous in situ wind measurements
between 50 and 60 km of altitude as provided by the
Pioneer Venus Night probe (Counselman et al. 1980)
and VEGA landers (Moroz & Zasova 1997). To reduce
the impact of transient phenomena like the equatorial
jets, the zonal winds obtained with cloud tracking are
presented as time averages of about 10 -- 20 days in most
of the cases, though this value strongly depends on the
data availability and how this is distributed along time.
Thus, zonal speeds from IRTF/SpeX images correspond
to nearly instantaneous winds (several hours) while the
average for 2004 May is provided by Limaye et al. (2006)
for observations performed along 70 days. The mean
zonal winds displayed in Fig. 13A -- I were obtained
from published results (Counselman et al. 1980; Allen
& Crawford 1984; Allen 1987; Moroz & Zasova 1997;
Crisp et al. 1989, 1991; Carlson et al. 1991; Chanover
et al. 1998; Limaye et al. 2006). The wind speeds from
2006 to 2017 correspond to a revisit of the VEx measure-
16
Peralta et al.
Figure 12. Winds' dependence with local time and longitude when the measurements from VEx/VIRTIS-M (2006 -- 2008) and
Akatsuki/IR2 (2016) are combined. Averages and standard deviation are calculated for bins of latitude, longitude and local
time of 10◦, 10◦ and 40' respectively. Panels (A -- D) display the zonal and meridional winds' dependence with the local time
and the longitude, while panels (E -- H) display the corresponding errors. Since the mean zonal winds during 2016 were faster
than during the VEx mission (see Fig. 5A), the zonal speeds used in panel (A) and (C) where multiplied by a correction factor
to have comparable zonally averaged profiles.
ments (Fig. 13J) performed by S´anchez-Lavega et al.
(2008); Hueso et al. (2012), and our measurements with
Akatsuki/IR2 (Fig. 13M), IRTF/SpeX (Fig. 13L,N)
and TNG/NICS (Fig. 13K).
Although caution must be taken when comparing val-
ues from different time-averages and images acquired
with different filters may provide some discrepancies
in the vertical level sensed, Fig. 13 suggests that the
zonal wind speeds at the nightside lower clouds seems
to experience long-term variation in the intensity of the
winds of up to 30 m s−1 (this total variation is larger
than variabilities linked by the local time or longitudinal
dependence described in subsections 4.2 and 4.3). The
zonal winds display a strong increase from the entry of
the Pioneer Venus probes in 1978 (Fig. 13A) and the
first cloud-tracked winds with ground-based observa-
tions (Fig. 13B) in the early 1980s. From 1983 to 1990
(Fig. 13B -- G) the zonal winds exhibit a gradual decrease
in magnitude, while from 2004 to 2008 the zonal winds
experience an increase above the error bars (Fig. 13J
from VEx data). This long-term behaviour of the night-
side zonal winds from VIRTIS-M images diverges from
the oscillation between 2007 and 2009 reported on the
zonal winds between 15◦S -- 25◦S of the dayside middle-
to-lower clouds found (Khatuntsev et al. 2017, fig. 11a
therein). The differences between the altitude range
sensed in day and night side images (S´anchez-Lavega
et al. 2008; Takagi & Iwagami 2011; Khatuntsev et al.
2017) or the distinct image sampling and time coverage
for the winds obtained from VMC and VIRTIS-M may
account for this discrepancy between VEx results.
The decadal variability for the zonal winds at the lower
clouds also keep similarities with the long-term trend
for the SO2 abundance at the cloud tops reported by
(Marcq et al. 2013) during 30 years of observations
combining Pioneer Venus and VEx observations. The
gradual decrease of the zonal winds along 1983 -- 1990
matches the decrease of the SO2 at the upper clouds
(Marcq et al. 2013, fig. 3 therein), what would support
the idea that the vertical level sensed by the nightside
images at 1.74, 2.26 and 2.32 µm might be variable.
This would make sense if the photochemical activity at
Nightside Winds of Venus with Akatsuki/IR2
17
Figure 13. Decadal variation of the zonal winds at the nightside lower clouds of Venus. Data correspond to time-averages
of the zonal winds obtained between 30◦S -- 30◦N with cloud tracking, except for (A) and (D) that represent instantaneous in
situ wind measurements from the Pioneer Venus Night probe (Counselman et al. 1980) and VEGA landers (Moroz & Zasova
1997) averaged within 50 -- 60 km of altitude. Blue dots represent time-averages using wind speeds provided in past publications:
(A) Counselman et al. (1980), (B) Allen & Crawford (1984), (C) Allen (1987), (D) Moroz & Zasova (1997), (E) Crisp et al.
(1989), (F) Crisp et al. (1991), (G) Carlson et al. (1991), (H) Chanover et al. (1998), (I) Limaye et al. (2006), and (J) Hueso
et al. (2012). New data presented in this work and based in cloud tracking measurements are displayed with red dots and were
obtained from TNG/NICS (K), IRTF/SpeX (L and N), and Akatsuki/IR2 images (M). The error bars stand for the standard
deviation of the time averages.
the upper clouds might could affect the total opacity of
the cloud layer (Knollenberg & Hunten 1980; Grinspoon
et al. 1993). In such a case, faster zonal speeds may be
coincident with periods when higher altitudes can be
sensed in the nightside images.
6. CONCLUSIONS
We have presented global results of the wind speeds at
the nightside lower clouds of Venus during the first year
of JAXA's Akatsuki mission. Both zonal and meridional
winds were obtained applying cloud tracking techniques
over 2.26-µm images acquired by the IR2 camera from
a selection of 466 images spanning the time interval
from 2016 March 22 to October 31. Automatic and
manual measurements were applied independently to
obtain 149,033 and 2,947 wind vectors, respectively.
In the specific case of the manual measurements, the
phase correlation technique was tested in the template
matching for the first time on Venus with comparable
results to those obtained using cross-correlation tech-
niques in the space domain. No dependence is found
between the wind speeds and the radiance or the size of
the cloud tracers. This supports recent reports of wind
variability due to strong horizontal shear and episodes
of jets at the equator, and the identification of clouds'
morphology consistent with shear instabilities. The
197819831988199319982003200820132018-40-50-60-70-80-90Zonal Wind (m s-1)ABCDEFGHIJKLMN18
Peralta et al.
meridional profiles of zonally-averaged wind speeds are
in overall agreement with results during 2006 -- 2008 from
Venus Express but are systematically 10 m s−1 faster.
Our results confirm in the northern hemisphere the
expected poleward decay of the zonal winds with sym-
metric winds between both hemispheres. Conversely
to past observations with Venus Express, zonal speeds
during the Akatsuki mission exhibits a local maximum
in the westward windspeeds caused by either a local
time dependence and/or influence of surface elevations,
although the irregular coverage of the dataset prevents a
definitive confirmation of its source. Also, a first analy-
sis of the decadal wind variability is performed for zonal
winds between 30◦S and 30◦N using a combination of
in situ and cloud tracking measurements performed be-
tween 1978 and 2017. Our results demonstrate yearly
and decade wind variability, suggesting that the zonal
winds at low latitudes might be affected by an oscillat-
ing disturbance with an amplitude of about ∼15 m s−1
and a period of about 30 years.
Finally,
in order to facilitate future studies of wind
variability and data comparison, we provide the full
data set of wind measurements performed with the
manual technique, accompanied by the template match
for individual measurements and animations of the pro-
jected image sequences used towards this purpose (see
accompanying Supplemental Material and Appendix A).
ACKNOWLEDGEMENTS
J.P. acknowledges JAXA's International Top Young
Fellowship.
R.H. and A.S.-L. were supported by
the Spanish MINECO project AYA2015-65041-P with
FEDER, UE support and Grupos Gobierno Vasco IT-
765-13. All authors acknowledge the work of the entire
Akatsuki team. We are also grateful to the anonymous
reviewer for his/her useful comments to improve the
manuscript.
APPENDIX
A. DESCRIPTION OF THE SUPPLEMENTAL MATERIAL
The supplemental material that accompanies this work consists on a compressed file that contains not only the
numerical values for our manual wind measurements obtained with the images of Akatsuki/IR2 camera, but also a
large set of images exhibiting the morphology of cloud tracers and the quality of the template matching, geometrical
projections of the IR2 images and animations of the cloud motions. The contents of the supplemental material are
described as it follows:
• a Readme file with a more detailed description of the contents of compressed file, as well as explanations about
the naming of files and folders.
• a Microsoft Excel file with the numerical values of the 2,277 manual wind measurements and where the motions
of the equatorial cloud discontinuity have been filtered out.
• a total of 103 folders for each pair of IR2 images, with wind measurements, animations and cloud tracers'
morphology and position.
REFERENCES
Acton, C. H. 1996, Planetary and Space Science, 44, 65
Ahmed, J., & Jafri, M. N. 2008, in Image and Signal
Bertaux, J.-L., Khatuntsev, I. V., Hauchecorne, A., et al.
2016, Journal of Geophysical Research (Planets), 121,
Processing, ed. A. Elmoataz, O. Lezoray, F. Nouboud, &
D. Mammass (Berlin, Heidelberg: Springer Berlin
Heidelberg), 128 -- 135
Allen, D. A. 1987, Icarus, 69, 221
Allen, D. A., & Crawford, J. W. 1984, Nature, 307, 222
Baffa, C., Comoretto, G., Gennari, S., et al. 2001,
Astronomy and Astrophysics, 378, 722
1087
Bevington, P. R., & Robinson, D. K. 1992, Data reduction
and error analysis for the physical sciences (New York:
McGraw-Hill, -- c1992, 2nd ed.)
Blamont, J. E., Young, R. E., Seiff, A., et al. 1986, Science,
231, 1422
Belton, M. J. S., Gierasch, P. J., Smith, M. D., et al. 1991,
Carlson, R. W., Baines, K. H., Kamp, L. W., et al. 1991,
Science, 253, 1531
Science, 253, 1541
Nightside Winds of Venus with Akatsuki/IR2
19
Chanover, N. J., Glenar, D. A., & Hillman, J. J. 1998,
Ikegawa, S., & Horinouchi, T. 2016, Icarus, 271, 98
Journal of Geophysical Research, 103, 31335
Iwagami, N., Sakanoi, T., Hashimoto, G. L., et al. 2018,
Counselman, C. C., Gourevitch, S. A., King, R. W., Loriot,
Earth, Planets, and Space, 70, 6
G. B., & Ginsberg, E. S. 1980, Journal of Geophysical
Research, 85, 8026
Crisp, D., Sinton, W. M., Hodapp, K.-W., et al. 1989,
Science, 246, 506
Crisp, D., McMuldroch, S., Stephens, S. K., et al. 1991,
Science, 253, 1538
del Genio, A. D., & Rossow, W. B. 1990, Journal of
Atmospheric Sciences, 47, 293
Fedorova, A., Marcq, E., Luginin, M., et al. 2016, Icarus,
275, 143
Folkner, W. M., Williams, J. G., & Boggs, D. H. 2009,
Interplanetary Network Progress Report, 178, 1
Foroosh, H., Zerubia, J. B., & Berthod, M. 2002, IEEE
Transactions on Image Processing, 11, 188
Fukuhara, T., Futaguchi, M., Hashimoto, G. L., et al. 2017,
Nature Geoscience, 10, 85
Gierasch, P. J., Goody, R. M., Young, R. E., et al. 1997,
The General Circulation of the Venus Atmosphere: an
Assessment, ed. S. W. Bougher, D. M. Hunten, & R. J.
Phillips (University of Arizona Press), 459
Gorinov, D. A., Khatuntsev, I. V., Zasova, L. V., Turin,
A. V., & Piccioni, G. 2018, Geophysical Research Letters,
45, 2554
Grinspoon, D. H., Pollack, J. B., Sitton, B. R., et al. 1993,
Planetary and Space Science, 41, 515
Hahn, G., & Jacquesson, M. 2012, WinJUPOS-Database
for object positions on planets and the Sun,
http://jupos.privat.t-online.de, ,
Horinouchi, T., Murakami, S.-y., Kouyama, T., et al. 2017a,
Measurement Science and Technology, 28, 085301
Horinouchi, T., Murakami, S., Satoh, T., et al. 2017b,
Nature Geoscience, 10, 646
Horinouchi, T., Kouyama, T., Lee, Y. J., et al. 2018, Earth,
Planets, and Space, 70, 10
Huang, H., Yoo, S., Yu, D., Huang, D., & Qin, H. 2012, in
Proceedings of the Twelfth International Workshop on
Multimedia Data Mining, MDMKDD '12 (New York,
NY, USA: ACM), 1 -- 9.
http://doi.acm.org/10.1145/2343862.2343863
Hueso, R., Peralta, J., Garate-Lopez, I., Bandos, T. V., &
S´anchez-Lavega, A. 2015, Planetary and Space Science,
113, 78
Hueso, R., Peralta, J., & S´anchez-Lavega, A. 2012, Icarus,
217, 585
Jun, L., & Fengxian, Z. 1992, Advances in Space Research,
12, 123
Khatuntsev, I. V., Patsaeva, M. V., Titov, D. V., et al.
2017, Journal of Geophysical Research (Planets), 122,
2312
-- . 2013, Icarus, 226, 140
Knollenberg, R. G., & Hunten, D. M. 1980, Journal of
Geophysical Research, 85, 8039
Kouyama, T., Imamura, T., Nakamura, M., Satoh, T., &
Futaana, Y. 2012, Planetary and Space Science, 60, 207
-- . 2013, Journal of Geophysical Research (Planets), 118,
37
Kouyama, T., Imamura, T., Taguchi, M., et al. 2017,
Geophysical Research Letters, 44, 12,098, 2017GL075792.
http://dx.doi.org/10.1002/2017GL075792
Kuglin, C. D. 1975, in Proc.International Conference on
Cyber-netics Society, 163 -- 165.
https://ci.nii.ac.jp/naid/20001697044/en/
Kuglin, C. D. 1976, Performance of the Phase Correlator in
Image Guidance Applications, Tech. rep., CONTROL
DATA CORP MINNEAPOLIS MN IMAGE SYSTEMS
DIV, Palo Alto Research Lab., Lockheed Missiles and
Space Company, Inc., Palo Alto, California
Lee, Y. J., Imamura, T., Schroder, S. E., & Marcq, E.
2015a, Icarus, 253, 1
Lee, Y. J., Titov, D. V., Ignatiev, N. I., et al. 2015b,
Planetary and Space Science, 113, 298
Lee, Y. J., Sawaga, H., Sato, T. M., et al. 2017a, in
Abstracts of JpGU-AGU Joint Meeting 2017, PPS06 -- P18
Lee, Y. J., Yamazaki, A., Imamura, T., et al. 2017b, The
Astronomical Journal, 154, 44
Leese, J. A., Novak, C. S., & Taylor, V. R. 1970, Pattern
Recognition, 2, 279
Limaye, S., Warell, J., Bhatt, B. C., Fry, P. M., & Young,
E. F. 2006, Bulletin of the Astronomical Society of India,
34, 189
Limaye, S. S. 2007, Journal of Geophysical Research
(Planets), 112, E04S09
Limaye, S. S., Watanabe, S., Yamazaki, A., et al. 2018,
Earth, Planets, and Space, 70, 38. https://link.
springer.com/article/10.1186/s40623-018-0789-5
Marcq, E., Bertaux, J.-L., Montmessin, F., & Belyaev, D.
2013, Nature Geoscience, 6, 25
Humblot, F., Collin, B., & Mohammad-Djafari, A. 2005, in
McGouldrick, K., Baines, K. H., Momary, T. W., &
Physics in Signal and Image Processing (PSIP)
conference, PSIP '05, Toulouse, France, 115 -- 120
Grinspoon, D. H. 2008, Journal of Geophysical Research
(Planets), 113, E00B14
20
Peralta et al.
McGouldrick, K., Momary, T. W., Baines, K. H., &
Grinspoon, D. H. 2012, Icarus, 217, 615
McGouldrick, K., & Toon, O. B. 2007, Icarus, 191, 1
-- . 2008, Icarus, 196, 35
McGouldrick, K., & Tsang, C. C. C. 2017, Icarus, 286, 118
Moroz, V. I., & Zasova, L. V. 1997, Advances in Space
Research, 19, 1191
Piccioni, G., Drossart, P., Suetta, E., et al., eds. 2007, ESA
Special Publication, Vol. 1295, VIRTIS: The Visible and
Infrared Thermal Imaging Spectrometer
Rayner, J. T., Toomey, D. W., Onaka, P. M., et al. 2003,
The Publications of the Astronomical Society of the
Pacific, 115, 362
Reddy, B. S., & Chatterji, B. N. 1996, IEEE Transactions
Mota Machado, P., Peralta, J., Luz, D., et al. 2016, in
on Image Processing, 5, 1266
AAS/Division for Planetary Sciences Meeting Abstracts,
Vol. 48, AAS/Division for Planetary Sciences Meeting
Abstracts, 115.06
Nakamura, M., Imamura, T., Ishii, N., et al. 2016, Earth,
Planets and Space, 68, 1.
http://dx.doi.org/10.1186/s40623-016-0457-6
Ogohara, K., Kouyama, T., Yamamoto, H., et al. 2012,
Icarus, 217, 661
Ogohara, K., Takagi, M., Murakami, S.-y., et al. 2017,
Earth, Planets, and Space, 69, 167
Parish, H. F., Schubert, G., Covey, C., et al. 2011, Icarus,
212, 42
Peralta, J., Hueso, R., & S´anchez-Lavega, A. 2007, Icarus,
190, 469
Peralta, J., Imamura, T., Read, P. L., et al. 2014, The
Astrophysical Journal Supplement Series, 213, 18.
http://stacks.iop.org/0067-0049/213/i=1/a=18
Peralta, J., Lee, Y. J., McGouldrick, K., et al. 2017a,
Icarus, 288, 235
Peralta, J., Luz, D., Berry, D. L., et al. 2012, Icarus, 220,
958
Peralta, J., Lee, Y. J., Hueso, R., et al. 2017b, Geophysical
Research Letters, 44, 3907, 2017GL072900.
http://dx.doi.org/10.1002/2017GL072900
Peralta, J., Hueso, R., S´anchez-Lavega, A., et al. 2017c,
Nature Astronomy, 1, 0187
Rossow, W. B., del Genio, A. D., & Eichler, T. 1990,
Journal of Atmospheric Sciences, 47, 2053
Samritjiarapon, O., & Chitsobhuk, O. 2008, in 2008
International Symposium on Communications and
Information Technologies, 364 -- 367
S´anchez-Lavega, A., Lebonnois, S., Imamura, T., Read, P.,
& Luz, D. 2017, Space Science Reviews, 212, 1541
S´anchez-Lavega, A., Hueso, R., Piccioni, G., et al. 2008,
Geophysical Research Letters, 35, 13204
Satoh, T., Nakamura, M., Ueno, M., et al. 2016, Earth,
Planets, and Space, 68, 74
Satoh, T., Sato, T. M., Nakamura, M., et al. 2017, Earth,
Planets, and Space, 69, 154
Schubert, G. 1983, General circulation and the dynamical
state of the Venus atmosphere, ed. D. M. Hunten,
L. Colin, T. M. Donahue, & V. I. Moroz (University of
Arizona Press), 681 -- 765
Takagi, M., Sugimoto, N., Ando, H., & Matsuda, Y. 2018,
Journal of Geophysical Research (Planets), 123, 335
Takagi, S., & Iwagami, N. 2011, Earth, Planets, and Space,
63, 435
Tavenner, T., Young, E. F., Bullock, M. A., Murphy, J., &
Coyote, S. 2008, Planetary and Space Science, 56, 1435
Titov, D. V., Svedhem, H., Koschny, D., et al. 2006,
Planetary and Space Science, 54, 1279
|
1004.1000 | 1 | 1004 | 2010-04-07T04:29:04 | Simulations for Terrestrial Planet Formation | [
"astro-ph.EP"
] | We investigate the formation of terrestrial planets in the late stage of planetary formation using two-planet model. At that time, the protostar has formed for about 3 Myr and the gas disk has dissipated. In the model, the perturbations from Jupiter and Saturn are considered. We also consider variations of the mass of outer planet, and the initial eccentricities and inclinations of embryos and planetesimals. Our results show that, terrestrial planets are formed in 50 Myr, and the accretion rate is about 60% - 80%. In each simulation, 3 - 4 terrestrial planets are formed inside "Jupiter" with masses of $0.15 - 3.6 M_{\oplus}$. In the 0.5 - 4AU, when the eccentricities of planetesimals are excited, planetesimals are able to accrete material from wide radial direction. The plenty of water material of the terrestrial planet in the Habitable Zone may be transferred from the farther places by this mechanism. Accretion may also happen a few times between two giant planets only if the outer planet has a moderate mass and the small terrestrial planet could survive at some resonances over time scale of $10^8$ yr. | astro-ph.EP | astro-ph | Icy Bodies of the Solar System
Proceedings IAU Symposium No. 263, 2009
Julio A. Fernandez,Daniela Lazzaro, & Dina Prialnik-Kovetz
c(cid:13) 2009 International Astronomical Union
DOI: 00.0000/X000000000000000X
Simulations for Terrestrial Planets
Formation
Jianghui JI1, and Niu ZHANG1,2,
1Purple Mountain Observatory, Chinese Academy of Sciences, Nanjing 210008, China
2Graduate School of Chinese Academy of Science, Beijing 100049
email: [email protected]
Abstract. We investigate the formation of terrestrial planets in the late stage of planetary
formation using two-planet model. At that time, the protostar has formed for about 3 Myr
and the gas disk has dissipated. In the model, the perturbations from Jupiter and Saturn are
considered. We also consider variations of the mass of outer planet, and the initial eccentricities
and inclinations of embryos and planetesimals. Our results show that, terrestrial planets are
formed in 50 Myr, and the accretion rate is about 60% - 80%. In each simulation, 3 - 4 terrestrial
planets are formed inside "Jupiter" with masses of 0.15 − 3.6M⊕. In the 0.5 - 4AU, when the
eccentricities of planetesimals are excited, planetesimals are able to accrete material from wide
radial direction. The plenty of water material of the terrestrial planet in the Habitable Zone
may be transferred from the farther places by this mechanism. Accretion may also happen a
few times between two giant planets only if the outer planet has a moderate mass and the small
terrestrial planet could survive at some resonances over time scale of 108 yr.
Keywords. methods:n-body simulations-planetary systems-planetary formation
1. Introduction
The discovery of the extrasolar planets (Mayor & Queloz 1995; Lee & Peale 2002;
Ji et al. 2003) around solar-type stars indeed provides substantial clues for the forma-
tion and origin of our own solar system. According to standard theory (Safronov 1969;
Wetherill 1990; Lissauer 1993), it is generally believed that planet formation may expe-
rience such several stages: in the early stage, the dust grains condense to grow km-sized
planetesimals; in the middle stage, Moon-to-Mars sized embryos are created by accretion
of planetesimals. When the embryos grow up to a core of ∼ 10M⊕, runaway accretion
may take place. With more gases accreted onto the solid core, the embryos become more
massive and eventually collapse to produce giant Jovian planets (Ida & Lin 2004). At
the end of the stage, it is around that the protostar has formed for about 3 Myr, the gas
disk has dissipated. A few larger bodies with low e and i are in crowds of planetesimals
with certain eccentricities and inclinations. In the late stage, the terrestrial embryos are
excited to high eccentricity orbits by mutual gravitational perturbation. Next, the orbital
crossings make planets obtain material in wider radial area. In this sense, solid residue
is either scattered out of the planetary system or accreted by the massive planet, even
being captured (Nagasawa & Ida 2000) at the resonance position of the giant planets.
Chambers (2001) made a study of terrestrial planet formation in the late stage by nu-
merical simulations, who set 150 − 160 Moon-to-Mars size planetary embryos in the area
of 0.3 − 2.0 AU under mutual interactions from Jupiter and Saturn. He also examined
two initial mass distributions: approximately uniform masses, and a bimodal mass distri-
bution. The results show that 2 − 4 planets are formed within 50 Myr, and finally survive
over 200 Myr timescale, and the final planets usually have eccentric orbits with higher
1
0
1
0
2
r
p
A
7
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
0
0
1
.
4
0
0
1
:
v
i
X
r
a
2
Ji & Zhang
eccentricities and inclinations . Raymond, Quinn & Lunine (2004, 2006) also investigated
the formation of terrestrial planets. In the simulations, they simply took into account
Jupiter's gravitational perturbation, and the distribution of material are in 0.5 − 4.5 AU.
Their results confirm a leading hypothesis for the origin of Earth's water: they may come
from the material in the outer area by impacts in the late stage of planet formation.
Raymond, Mandell, & Sigurdsson (2006) explored the planet formation under planetary
migration of the giant. In the simulations, super Hot Earth form interior to the migrating
giant planet, and water-rich, Earth-size terrestrial planet are present in the Habitable
Zone (0.8 − 1.5 AU) and can survive over 108 yr timescale.
In our work, we consider two-planet model, in which Jupiter and Saturn are supposed
to be already formed, with two swarms of planetesimals distributed in the region among
0.5 − 4.2 AU and 6.2 − 9.6 AU respectively. The initial eccentricities and inclinations of
planetesimals are considered. We also vary the mass of Saturn to examine how the small
bodies evolve. The simulations are performed on longer timescale 400 Myr in order to
check the stability and the dynamical structure evolution of the system. In the following,
we briefly summarize our numerical setup and results .
2. Numerical Setup
The timescale for formation of Jupiter-like planets is usually considered to be less
than 10 Myr (Briceno 2001), the formation scenario of planet embryos is related to
their heliocentric distances and the initial mass of the star nebular. If we adopt the
model of 1.5 MMSN (Minimum Mass Solar Nebular), the upper bound of the timescale
for Jupiter-like planet formation corresponds to the timescale for embryo formation at
2.5 AU (Kokubo & Ida 2002), which is just at 3 : 1 resonance location of Jupiter. In
the region 2.5 − 4.2 AU, embryos will be cleared off by strong perturbation from Jupiter.
There should be some much smaller solid residue among Jupiter and Saturn, even though
the clearing effect may throw out most of the material in this area. We set embryos simply
in the region 0.5 − 2.5 AU and planetesimals at 0.5 − 4.2 AU and 6.2 − 9.6 AU.
We adopt the surface density profile as follows (Raymond, Quinn & Lunine 2004):
Σ(r) = (cid:26) Σ1r−3/2,
Σsnow( r
5AU )−3/2,
r < snow line,
r > snow line.
(2.1)
In (2.1), Σsnow = 4 g/cm2 is the surface density at snowline, where the snowline is
at 2.5 AU with Σ1 = 10 g/cm2. The mass of planetary embryos is proportional to the
width of the feeding zone, which is associated with Hill Radius, RH , so the mass of an
embryo increases as
The embryos in the 0.5 − 2.5 AU are spaced by Λ (Λ varying randomly between 2 and
5) mutual Hill Radii, RH,m, which is defined as
Membryo ∝ rΣ(r)RH
(2.2)
RH,m = (
a1 + a2
2
)(
m1 + m2
3M⊙
)1/3
(2.3)
where a1,2 and m1,2 are the semi-major axes and masses of the embryos respectively.
Replacing RH in (2.2) with RH,m, and substituting (2.1) in (2.2), then, we achieve a
relation law between the mass of embryos and the parameter Λ as
Membryo ∝ rΣ(r)RH,m ∝ r3/4Λ3/2Σ3/2
(2.4)
Here, we equally set the masses of planetesimals inside and outside Jupiter, respec-
Terrestrial Planets Formation
3
e
e
e
0.0 0.1 0.2 0.3 0.4
0.0 0.1 0.2 0.3 0.4
0.0 0.1 0.2 0.3 0.4
0 Myr
1
/
3
1
/
2
2
/
3
0.1 Myr
1 Myr
10 Myr
50 Myr
400 Myr
1 2 3 4 5 6 7 8 9
1 2 3 4 5 6 7 8 9
a(AU)
a(AU)
e
e
e
0.0 0.1 0.2 0.3 0.4
0.0 0.1 0.2 0.3 0.4
0.0 0.1 0.2 0.3 0.4
0 Myr
1
3
/
1
2
/
2
3
/
0.1 Myr
1 Myr
10 Myr
50 Myr
400 Myr
3
2
/
2
1
/
1 2 3 4 5 6 7 8 9
1 2 3 4 5 6 7 8 9
a(AU)
a(AU)
Log(Water Mass Fraction)
Log(Water Mass Fraction)
−5
−4
−3
−2
−1.3
−5
−4
−3
−2
−1.3
Figure 1. Left panel : (a) Snapshot of simulation 2a with MSaturn = 5M⊕. The total mass of
embryos is 2.4M⊕, the masses of planetesimals inside Jupiter are 0.0317M⊕, and those outside
Jupiter are 0.0375M⊕. Planetesimals among Jupiter and Saturn were nonself-gravitational (see
Section 2.1). Note the size of each object is relative, and the value bar is log of water mass
fraction. Right panel : (b) snapshot for similuation 2b.
tively. Consequently, the number distribution of the planetesimals is simply required to
meet N ∝ r−1/2. Additionally, we remain the total number of planetesimals and embryos
inside Jupiter, and the number of planetesimals outside Jupiter both equal to 200. The
mass inside and outside Jupiter is equal to be 7.5M⊕. The eccentricities and inclinations
vary in (0 − 0.02) and (0 − 0.05◦), respectively. The mass of Saturn in simulations 1a/1b,
2a/2b and 3a/3b are 0.5M⊕, 5M⊕, 50M⊕ respectively. Each simulations marked by label
a (or b) is run to consider (not consider) self-gravitation of planetesimals among giants.
We use the hybrid symplectic integrator (Chambers 1999) in MERCURY package to
integrate all the simulations. In addition, we adopt 6 days as the length of time step,
which is a twentieth period of the innermost body at 0.5 AU. All runs are carried out
over 400 Myr time scale. At the end of the intergration, the changes of energy and
angular momenta are 10−3 and 10−11 respectively. Six simulations are performed on a
workstation composed of 12 CPUs with 1.2 GHz, and each costs roughly 45 days.
3. Results
Fig. 1(a) is a snapshot of simulation 2a. At 0.1 Myr, it is clear that the planetesimals are
excited at the 3 : 2 (3.97 AU),2 : 1 (3.28 AU) and 3 : 1 (2.5 AU) resonance positions with
Jupiter, and this is quite similar to the Kirkwood gaps of the main asteroidal belt in solar
system. For about 1 Myr, planetesimals and embryos are deeply intermixed, where most
of the bodies have stirred to be large eccentricities. Collisions and accretions frequently
emerge among planetesimals and embryos. This process continues until about 50 Myr,
and the planetary embryos are mostly generated. The formation timescale of embryos is
in accordance with that of (Ida & Lin 2004). Finally, inside Jupiter, 3 terrestrial planets
are formed with masses of 0.15 − 3.6M⊕. However, at the outer region, the planetesimals
are continuously scattered out of the system at 0.1 Myr. For about 10 Myr, there are
no survivals except at some resonances with the giant planet. As shown in the Figure,
there is a small body at the 1 : 2 resonance with Jupiter. Due to scattering amongst
planetesimals, Jupiter (Saturn) migrates inward (outward) 0.13 AU (1.19 AU) toward
the sun respectively. Such kind of migration agrees with the work of Fernandez et al.
(Fernandez & Ip 1984) Hence, the 2 : 5 mean motion resonance is destroyed, then the
ratios of periods between Jupiter and Saturn degenerate to 1 : 3. Therefore, the ratio of
4
Ji & Zhang
Table 1. Properties of terrestrial planets from different systems
¯i(◦)
System accretion rate n ¯m(m⊕) concentration
¯e
1a
1b
2a
2b
3a
3b
1a-3b
solar
3
3
3
4
3
3
1.8313
73.2518%
2.0096
80.3853%
1.4958
59.8322%
1.3683
72.9779%
1.6277
65.1098%
66.9694%
1.6742
69.7544% 3.2 1.6678
0.4943
-
4
0.4606
0.4262
0.8116
0.4299
0.5337
0.5040
0.5276
0.5058
0.1381 7.6963
0.0937 1.7790
0.2108 16.9117
0.0999 5.1415
0.2063 5.9153
0.1839 5.2447
0.1554 7.1148
0.0764 3.0624
periods for Jupiter, small body and Saturn is approximate to 1 : 2 : 3. In the 0.5 − 4 AU,
when the eccentricities of planetesimals are excited, planetesimals are able to accrete
material from wide radial direction. The plenty of water material of the terrestrial planet
in the Habitable Zone may be transferred from the farther places by this mechanism.
Fig. 1(b) is illustrated for simulation 2b. In comparison with Fig. 1(a), it is apparent
that planetesimals are excited more quickly at the 3 : 2 (3.97 AU), 2 : 1 (3.28 AU) and
3 : 1 (2.5 AU) resonance location with Jupiter. The several characteristic timescales are
the same as simulation 2a for the bodies within Jupiter. 4 planets are formed in simulation
2b, the changes of position of Jupiter and Saturn behaves like simulation 2a. We point
out that simulations 2a and 2b share the initial conditions, but the only difference in
them is whether we consider the self-gravitation among the outer planetesimals. There
is a little gathered planetesimals survival over 400 Myr among 7 − 8 AU, located in the
area of 2 : 3 (6.63 AU) and 1 : 2 (8.03 AU) resonances with Jupiter. The detailed results
for whole simulations that the reader may refer to (Zhang & Ji 2009).
The production efficiency of the terrestrial planet in our model is high, and the accre-
tion rate inside Jupiter is 60% − 80% in the simulations. 3 − 4 terrestrial planets formed
in 50 Myr. 5 of 6 simulations have a terrestrial planet in the Habitable Zone (0.8 − 1.5
AU). The planetary systems are formed to have nearly circular orbit and coplanarity,
similar to the solar system (see Table 1). We suppose that the above characteristics are
correlated with the initial small eccentricities and inclinations. The concentration in Ta-
ble 1 means the ratio of maximum terrestrial planet formed in the simulation and the
total terrestrial planets mass. It represents different capability on accretion. The average
value of this parameter is similar to the solar system. Considering the self-gravitation
of planetesimals among Jupiter and Saturn, the system has a better viscosity, so that
the planetesimals will be excited slower. The consideration of self-gravitation may not
change the formation time scale of terrestrial planets, but will affect the initial accretion
speed and the eventual accretion rate.
4. Summary and Discussion
We simulate the terrestrial planets formation by using two-planet model. In the simula-
tion, the variations of the mass of outer planet, the initial eccentricities and inclinations
of embryos and planetesimals are also considered. The results show that, during the
terrestrial planets formation, planets can accrete material from different regions inside
Jupiter. Among 0.5 − 4.2 AU, the accretion rate of terrestrial planet is 60% − 80%, i.e.,
about 20% − 40% initial mass is removed during the progress. The planetesimals will
improve the efficiency of accretion rate for certain initial eccentricities and inclinations,
and this also makes the newly-born terrestrial planets have lower orbital eccentricities.
Terrestrial Planets Formation
5
It is maybe a common phenomenon in the planet formation that the water-rich terres-
trial planet is formed in the Habitable Zone. The structure, which is similar to that of
solar system, may explain the results of disintegration of a terrestrial planet. Most of the
planetesimals among Jupiter and Saturn are scattered out of the planetary systems, and
this migration caused by scattering (Fernandez & Ip 1984) or long-term orbital evolution
can make planets capture at some mean motion resonance location. Accretion could also
happen a few times between two planets if the outer planet has a moderate mass, and the
small terrestrial planet could survive at some resonances over 108 yr time scale. Struc-
turally, Saturn has little effect on the architecture inside Jupiter, owing to its protection.
However, obviously, a different Saturn mass could play a vital role of the structure outer
Jupiter. Jupiter and Saturn in the solar system may form over the same period.
Acknowledgements
This work is financially supported by the National Natural Science Foundations of
China (Grants 10973044, 10833001, 10573040, 10673006, ) and the Foundation of Minor
Planets of Purple Mountain Observatory.
References
Briceno, C. et al., 2001, Science, 291, 93
Chambers, J. E. 1999, MNRAS, 304, 793
Chambers, J. E. 2001, Icarus, 152, 205
Fernandez, J. A., & Ip, W. H. 1984, Icarus, 58, 109
Ida, S., & Lin, D. N. C. 2004, ApJ, 604, 388
Ji, J. H., et al.2003, ApJ, 585, L139
Kokubo, E., & Ida, S. 2002, ApJ, 581, 666
Lee, M. H., & Peale, S. J. 2002, ApJ, 567, 596
Lissauer, J.J. 1993, ARAA, 31, 129
Nagasawa, M., & Ida, S. 2000,AJ, 120, 3311
Raymond, S. N., Quinn, T., & Lunine, J. I. 2004, Icarus, 168, 1
Raymond, S. N., Quinn, T., & Lunine, J. I. 2006, Icarus, 183, 265
Raymond, S. N., Mandell, A. M., & Sigurdsson, S. 2006, Science, 313, 1413
Safronov,V.S. 1969, Evolution of the Protoplanetary Cloud and Formation of the Earth and the
Planets (Moscow:Nauka)
Mayor, M., & Queloz, D. 1995, Nature, 378, 355
Wetherill, G. W. 1990, Ann. Rev. Earth Planet Sci., 18, 205
Zhang N., Ji J., 2009, Science in China Series G, 52(5), 794
|
1711.05185 | 1 | 1711 | 2017-11-14T16:44:41 | ALMA continuum observations of the protoplanetary disk AS 209. Evidence of multiple gaps opened by a single planet | [
"astro-ph.EP",
"astro-ph.SR"
] | The paper presents new high angular resolution ALMA 1.3 mm dust continuum observations of the protoplanetary system AS 209 in the Ophiuchus star forming region. The dust continuum emission is characterized by a main central core and two prominent rings at $r = 75\,$au and $r = 130\,$au intervaled by two gaps at at $r = 62\,$au and $r = 103\,$au. The two gaps have different widths and depths, with the inner one being narrower and shallower. We determined the surface density of the millimeter dust grains using the 3D radiative transfer disk code \textsc{dali}. According to our fiducial model the inner gap is partially filled with millimeter grains while the outer gap is largely devoid of dust. The inferred surface density is compared to 3D hydrodynamical simulations (FARGO-3D) of planet-disk interaction. The outer dust gap is consistent with the presence of a giant planet ($M_{\rm planet} \sim 0.8\,M_{\rm Staturn}$); the planet is responsible for the gap opening and for the pile-up of dust at the outer edge of the planet orbit. The simulations also show that the same planet can give origin to the inner gap at $r = 62\,$au. The relative position of the two dust gaps is close to the 2:1 resonance and we have investigated the possibility of a second planet inside the inner gap. The resulting surface density (including location, width and depth of the two dust gaps) are in agreement with the observations. The properties of the inner gap pose a strong constraint to the mass of the inner planet ($M_{\rm planet} < 0.1\,M_{\rm J}$). In both scenarios (single or pair of planets), the hydrodynamical simulations suggest a very low disk viscosity ($\alpha < 10^{-4}$). Given the young age of the system (0.5 - 1 Myr), this result implies that the formation of giant planets occurs on a timescale of $\lesssim$ 1\,Myr. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. as209
November 15, 2017
c(cid:13)ESO 2017
ALMA continuum observations of the protoplanetary disk AS 209
Evidence of multiple gaps opened by a single planet
D. Fedele1, M. Tazzari2, R. Booth2, L. Testi3, C. J. Clarke2,
I. Pascucci4, A. Kospal5, D. Semenov6, S. Bruderer, Th. Henning6, and R. Teague6
1 INAF-Osservatorio Astrofisico di Arcetri, L.go E. Fermi 5, I-50125 Firenze, Italy
2 Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
3 European Southern Observatory, Karl-Schwarzschild-Strasse 2, 85748, Garching bei Muenchen, Germany
4 Lunar and Planetary Laboratory, The University of Arizona, Tucson, AZ 85721, USA
5 Konkoly Observatory, Research Centre for Astronomy and Earth Sciences, Hungarian Academy of Sciences, Konkoly-Thege Mik-
lós út 15-17, 1121 Budapest, Hungary
6 Max Planck Institute for Astronomy, Königstuhl 17, 69117, Heidelberg, Germany
Received ...; accepted ...
ABSTRACT
The paper presents new high angular resolution ALMA 1.3 mm dust continuum observations of the protoplanetary system AS 209 in
the Ophiuchus star forming region. The dust continuum emission is characterized by a main central core and two prominent rings at
r = 75 au and r = 130 au intervaled by two gaps at at r = 62 au and r = 103 au. The two gaps have different widths and depths, with
the inner one being narrower and shallower. We determined the surface density of the millimeter dust grains using the 3D radiative
transfer disk code dali. According to our fiducial model the inner gap is partially filled with millimeter grains while the outer gap
is largely devoid of dust. The inferred surface density is compared to 3D hydrodynamical simulations (FARGO-3D) of planet-disk
interaction. The outer dust gap is consistent with the presence of a giant planet (Mplanet ∼ 0.8 MStaturn); the planet is responsible for
the gap opening and for the pile-up of dust at the outer edge of the planet orbit. The simulations also show that the same planet
can give origin to the inner gap at r = 62 au. The relative position of the two dust gaps is close to the 2:1 resonance and we have
investigated the possibility of a second planet inside the inner gap. The resulting surface density (including location, width and depth
of the two dust gaps) are in agreement with the observations. The properties of the inner gap pose a strong constraint to the mass of
the inner planet (Mplanet < 0.1 MJ). In both scenarios (single or pair of planets), the hydrodynamical simulations suggest a very low
disk viscosity (α < 10−4). Given the young age of the system (0.5 - 1 Myr), this result implies that the formation of giant planets
occurs on a timescale of (cid:46) 1 Myr.
Key words. giant planet formation – T Tauri
1. Introduction
Axisymmetric gaps and rings such as those seen in the pro-
toplanetary disks around HL Tau, TW Hya, HD163296, HD
169142, AA Tau (e.g. ALMA Partnership et al. 2015; Andrews
et al. 2016; Isella et al. 2016; Fedele et al. 2017; Loomis et al.
2017) can now regularly be unveiled by the extremely high res-
olution available with ALMA. The formation of gaps and rings
in disks can be due to several mechanisms such as: planet for-
mation (e.g., Papaloizou & Lin 1984); magneto-rotational in-
stability (Flock et al. 2015); condensation fronts (Zhang et al.
2015); dust sintering (Okuzumi et al. 2016); photoevaporation
(Ercolano et al. 2017).
The rings observed by ALMA show that the millimeter dust
grains are radially confined regions in which inward radial mi-
gration is slowed down or stopped and so may be key to explain-
ing the retention of large grains in disks on long (2 − 3 Myr)
timescales (irrespective of the mechanism producing such "dust
traps"). In addition, dust traps provide the ideal environment in
which to observationally constrain models of grain growth be-
cause – in contrast to other regions of the disk – they are re-
gions in which radial drift is relatively unimportant, allowing
dust to grow in situ (Pinilla et al. 2012). Given a measure of
the local dust density, the timescale for dust growth to a given
size is readily obtained from grain growth models (e.g., Birn-
stiel et al. 2012). Empirical measurement of the maximum grain
size in traps would thus provide the cleanest test of the various
assumptions (sticking probability, turbulent velocity, fragmenta-
tion threshold) that enter these models.
This paper presents new ALMA 1.3 mm continuum observa-
tions of the T Tauri disk AS 209 where axisymmetric gaps and
rings are detected. AS 209 (M(cid:63) = 0.9 M(cid:12), spectral type K5,
L(cid:63) = 1.5 L(cid:12), Tazzari et al. 2016) is part of the young (age
∼ 0.5 − 1.0 Myr Natta et al. 2006) Ophiuchus star forming re-
gion at a distance of 126 pc from the Sun (Gaia Collaboration
et al. 2016). Multi-frequency continuum observations revealed
Article number, page 1 of 9
7
1
0
2
v
o
N
4
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
8
1
5
0
.
1
1
7
1
:
v
i
X
r
a
A&A proofs: manuscript no. as209
Fig. 1. ALMA 1.3 mm dust continuum image (uniform weighting). The main substructures are highlighted in the right panel.
optically thin emission at millimeter wavelenghts beyond a few
10s of au from the star (Pérez et al. 2012; Tazzari et al. 2016).
Huang et al. (2016) found evidence of an extended gas emis-
sion (C18O) speculating that it is due to external CO desorption
in the outer disk. Interestingly, Huang et al. (2017) noticed the
presence of a dark lane in the dust 1.1 mm continuum emission.
The structure of the paper is the following: observations and data
reduction are presented in section 2 and the results are discussed
in section 3. The data analysis is described in Section 4. Sec-
tion 5 provides a comparison to hydrodynamical simulations.
Discussion and conclusion are reported in sec. 6.
2. Observations and Data reduction
The ALMA observations of AS 209 (J2000: R.A. =
16h49m15.296s, DEC = –14◦22(cid:48)09.02(cid:48)(cid:48)) have been performed on
2016 September 22 (with 38 antennas) and 26 (41 antennas) in
band 6 (211–275 GHz) as part of the project ID 2015.1.00486.S
(PI: D. Fedele). The correlator setup includes a broad (2 GHz
bandwidth) spectral window centered at 230 GHz.
Visibilities were taken in two execution blocks with a 6.05s in-
tegration time per visibility totalling 40 minutes, per block, on-
source. System temperatures were between 80− 145 K. Weather
conditions on the dates of observation gave an average precip-
itable water vapour of 2.2 and 2.3 mm, respectively. Calibration
was done with J1517−2422 as bandpass calibrator, J1733−1304
as phase and flux the flux calibrator. The visibilities were subse-
quently time binned to 60s integration times per visibility for
self-calibration, imaging, and analysis. Self-calibration was per-
formed using the 233 GHz continuum TDM spectral window
with DA41 as the reference antenna.
The continuum image was created using casa.clean (casa ver-
sion 4.7); after trying different weighting schemes, we opted for
a uniform weight which yields a synthesised beam of 0(cid:48)(cid:48).19 ×
0(cid:48)(cid:48).14 (PA = 75.5◦). The peak flux is 13 mJy beam−1 and the r.m.s.
is 0.1 mJy beam−1.
Article number, page 2 of 9
3. Results
The ALMA 1.3 mm dust continuum image is shown in Fig. 1:
the continuum emission is characterized by a bright central emis-
sion and two weaker dust rings peaking at ∼ 75 au and 130 au,
respectively. The two rings have a similar peak flux (∼ 2 mJy).
The rings are intervaled by two narrow gaps. The two gaps have
different widths and depths. The radial intensity profile shows a
kink around 20 − 30 au which may be the signature of another
(spatially unresolved) dust gap. Finally, the continuum flux does
not drop to zero at the edge of the outer ring as there is a tenuous
emission extending out to ∼ 170−180 au. The different disk sub-
structures are clearly visible in the radial intensity profile shown
in Fig. 2.
3.1. Characterization of the brightness profile
In this section we present here the fit of the observed visibil-
ities, which provides an initial characterization of the disk sur-
face brightness useful for the detailed physical modelling carried
out in Section 4. We assume an axisymmetric brigthness profile
defined as follows:
(cid:33)φ1
(cid:32) R
Rc
−
(cid:32) R
Rc
(cid:33)φ2 ,
I(R) = δ(R) I0
exp
where I0 is a normalization, Rc is a scale length and δ(R) is a
scaling factor (by definition δ(R) > 0) parametrized as:
for R ∈ [RG1 − hwG1, RG1 + hwG1]
for R ∈ [RG1 + hwG1, RG2 − hwG2]
for R ∈ [RG2 − hwG2, RG2 + hwG2]
for R ∈ [RG2 + hwG2, RR2,out]
for R ≥ RR2,out
otherwise
δG1
δR1
δG2
δR2
δout
1
δ(R) =
(1)
(2)
D. Fedele et al.: ALMA continuum observations of the protoplanetary disk AS 209
Fig. 2. Radial intensity profile of the 1.3, mm dust continuum emission.
The profile is azimuthally averaged after deprojecting for the disk incli-
nation (i = 36◦, Sect. 3). The black line shows the mean profile while
the shadowed regions indicate the standard deviation along the azimuth
angle.
Table 1. Parameter space explored by the Markov chains and best-fit
values
Parameter
I0 [mJy/beam]
Rc [au]
φ1
φ2
RG1 [au]
hwG1 [au]
δG1
δR1
RG2 [au]
hwG2 [au]
δG2
δR2
RR2,out [au]
δout
i [◦]
PA [◦]
Min Max
100
150
4
4
80
30
1
3
110
30
1
20
180
2
90
180
0
20
-4
-4
0
0
0
0
80
0
0
0
130
0
0
0
Best-fit
7.4 ± 0.2
80 ± 1
-0.24 ± 0.01
2.19 ± 0.02
61.7 ± 0.5
8.0 ± 0.2
0.03 ± 0.005
0.80 ± 0.02
103.2 ± 0.4
15.6 ± 0.2
0.025 ± 0.005
4.8 ± 0.1
139.8 ± 0.8
1.95 ± 0.03
35.3 ± 0.8
86.0 ± 0.7
where RG and hwG are the center and half width of the dust
gaps, respectively. The choice of this particular brightness pro-
file serves as a simple realization of an "unperturbed" profile (an
exponentially tapered power law), characterized by a few radial
regions that can depart from it either due to an excess (δ > 1)
or a lack (δ < 1) of emission. Following the evidence emerging
from the synthesized image (Fig. 1), we allow for two rings, two
gaps, and an outer disk region, following nomenclature in Fig. 1.
In this framework, a gap in the disk is naturally modelled with
δ < 1.
We perform the fit of the visibilities with a Bayesian ap-
proach using the Monte Carlo Markov chains ensemble sampler
Fig. 3. Dust surface density of the dali fiducial model. The dashed line
corresponds to the initial unperturbed profile (power-law with exponen-
tial tail).
(Goodman & Weare 2010) implemented in the emcee package
(Foreman-Mackey et al. 2013). We assume flat priors for all the
free parameters. Table 1 reports the ranges explored for each pa-
rameter.
We also fit simultaneously for the disk inclination i, the East-
of-North position angle PA, and phase center offset (∆α, ∆δ).
For each set of values of the free parameters we compute a syn-
thetic image of the model assuming axisimmetry and we use the
galario library (Tazzari et al. 2017) to compute the synthetic vis-
ibilities (by sampling the Fourier transform of the model image
in all the observed (u, v)-points) and the resulting χ2 as
N(cid:88)
j=0
χ2 =
Vobs(u j, v j) − Vmod(u j, v j)2w j ,
(3)
where w j is the weight of the observed (u j, v j) visibility point.
The posterior of each model is then computed as exp(−χ2/2)
and sampled with 80 chains for 45000 steps (after 5000 burn-
in steps). The chains, which reached a good convergence, are
shown in Fig. A.2 in Appendix A, in the form of marginalized 1D
and 2D distributions. The parameters that we infer from the fit of
the visibilities are presented in Table 1: for each parameter, we
estimate its value as the median of the marginalized distribution
and its uncertainty as half the interval between 16% and 84%
percentiles.
We find that the 1.3 mm continuum brightness distribution of
AS 209 can be explained very well by a profile with two deep
gaps (δG1 ∼ 0.03, δG2 ∼ 0.025) at 62 and 103 au, respectively,
and an excess ring (δR2 ∼ 4) at ∼ 130 au. The excellent agree-
ment of this profile with the observations is apparent in Fig. A.1,
where the deprojected synthetic visibilities match the observed
ones up to 1500kλ.
Article number, page 3 of 9
Table 2. dali fiducial disk model
A&A proofs: manuscript no. as209
Fixed
Parameter
M(cid:63) [M(cid:12)]
Teff [K]
L(cid:63) [L(cid:12)]
d [pc]
Rin [au]
Rc [au]
RG1 [au]
hwG1 [au]
RG2 [au]
hwG [au]
RR2,out [au]
Rout [au]
i [◦]
PA [◦]
χ, flarge
ψ
hc
Variable
Parameter Min
Mdust [M(cid:12)]
γ1
γ2
δG1
δR1
δG2
δR2
δout
Value
0.9†
4250†
1.5†
126
0.1
80
62
8
103
16
140
200
35
86
0.2, 0.85
0.1†
0.133†
Description
stellar mass
stellar temperature
stellar luminosity
stellar distance
disk inner radius
disk critical radius
Gap 1 center
Gap 1 half width
Gap 2 center
Gap 2 half width
Ring 2 outer radius
disk outer radius
disk inclination
disk position angle
settling parameters
flaring exponent
scale height at Rc
Max
5 · 10−4
1.0
3.0
Step
0.5 · 10−4
0.1
0.1
1 · 10−4
-1.0
1.0
(0, 0.001, 0.01, 0.05, 0.10, 0.15, 0.20)
0.5
(0, 0.001, 0.01, 0.02, 0.03, 0.05, 0.1)
2
1
0.5
0.5
0.05
1.0
5
3
Fiducial
3.5 · 10−4 Disk dust mass
0.3
2.0
0.1
0.75
0.01
4.5
1.5
Σ(r) power-law exponent
Σ(r) exponential-tail exponent
Σdust,large scale factor in gap 1
Σdust,large scale factor in ring 1
Σdust,large scale factor in gap 2
Σdust,large scale factor in ring 2
Σdust,large scale factor in outer disk
Notes. For each variable parameter we explored a range of values between a minimum (Min) and maximum (Max) in regular steps (step). In the
case of δG1 and δG2 we explored non-linearly spaced values, so we list them in parenthesis. References: (†) Andrews et al. (2009) .
4. Analysis with a physical disk model
In this Section we aim to characterize the structure of AS 209
in physical terms, starting from the observed 1.3mm continuum
observation. This step is important to estimate the drop of the
dust surface density inside the two gaps. For this purpose we use
the dust radiative transfer code implemented into the thermo-
chemical disk model dali (Dust and Lines, Bruderer et al. 2012;
Bruderer 2013). Starting from an input radiation field and from a
disk density structure, dali solves the two dimensional dust con-
tinuum radiative transfer and determines the dust temperature
and radiation field strength at each disk position.
4.1. Model description
We adopt the characterization of the surface brightness presented
in the previous Section as a first guess for the functional form and
the location of the gaps to be used for the surface density of the
physical disk model that we use in this Section. The dust surface
density is:
(cid:33)γ1
(cid:32) R
Rc
(cid:34)
−
exp
(cid:33)γ2(cid:35)
(cid:32) R
Rc
Σdust(R) = δ(R) Σc
(4)
where the surface density scaling factor (δ) is parametrized as in
eq. 3.1.
Article number, page 4 of 9
(cid:33)ψ
(cid:32) R
Rc
h = hc
In the vertical direction, the density follows a Gaussian distribu-
tion with scale height h (= H/R)
(5)
with hc the critical scale height and ψ the flaring exponent. The
critical scale height (hc = 0.13) and disk flaring (ψ = 0.1) are
taken from Andrews et al. (2011). In the adopted version of dali,
dust settling is included following D'Alessio et al. (2006), i.e.,
adopting two power-law grain size populations with different
scale heights: the small grains have a scale height equal to h
(similar to the gas) while the scale height of the large grains is
χh (with χ < 1) to account for the settling of the large grains.
Finally, the total dust mass is distributed between the two popu-
lations and it is regulated by the parameter flarge (large-to-small
mass ratio): thus, the dust surface density is Σdust · (1 − flarge)
and Σdust · flarge for the small and large grains, respectively. The
flaring parameters are fixed : χ = 0.2, flarge = 0.85.
We fixed the grain size populations with a small population
of sizes between 0.005 and 1 µm and a large one of sizes be-
tween 0.005 and smax µm with both populations sharing the same
power-law exponent (p = 3.5). The dust mass absorption coeffi-
cients are taken from Andrews et al. (2011).
The maximum grain size of the large dust population affects the
dust opacity at millimeter wavelengths with the opacity decreas-
ing by almost an order of magnitude going from smax = 0.8 mm
D. Fedele et al.: ALMA continuum observations of the protoplanetary disk AS 209
Fig. 4. dali fiducial model, comparison of 1.3 mm continuum image. The model image is produced with casa.clean starting from the synthetic
visibilities and adopting the same clean parameters as the observations.
to 1.0 cm (e.g., Tazzari et al. 2016). This in turn has an impact
the dust temperature and the total dust mass. In this paper we fix
smax = 2000 µm in agreement with the multi-frequency contin-
uum analysis of AS 209 by Tazzari et al. (2016).
4.2. Fiducial model
We built a grid of dali disk models varying the following param-
eters:
• total dust mass
• surface density power-law exponent (γ1)
• exponential tail exponent (γ2)
• dust density scaling factors
The explored range of each variable parameter is listed in Ta-
ble 2. The values of Rc, gaps centers (RG1, RG2) and gaps sizes
(half width hwG1, hwG2) are taken from the best-fist results of the
multi-components analysis.
Each model is compared with the ALMA continuum observa-
tion with the aim of defining a fiducial model for the dust sur-
face density. The comparison is performed in the uv-plane: first
we compute the synthetic 1.3 mm continuum image with dali,
then the tool casa.simobserve is used to convert the image into
synthetic visibilities at the same uv-positions as the observations.
Finally we measure the χ2 between observation and model after
deprojecting and binning the visibilities (bin size of 30 kλ).
The fitting procedure is performed in multiple steps: first we vary
the global disk properties, i.e. Mdust, γ1, γ2 until we find a good
agreement with the visibilities at the shortest baselines which
provides a constraint to the large scale structure. During this step
the dust scaling factors are kept fixed: δG1 = 0, δG2 = 0, δR1 =
1, δR2 = 1, δout = 1. In a second step, we constrain the values of
the scaling factors while keeping fixed Mdust, γ1, γ2. The process
is repeated until convergence. This allows us to refine the grid
resolution iteratively.
The fiducial model is defined by the set of parameters that min-
imize the χ2 between the observed and synthetic visibilities
(eq. 3.1) within the explored parameter space.
The parameters of the dali fiducial model are listed in Table 2.
Figure 3 shows the dust surface density of the fiducial model and
the model image is shown in Figure 4. We note that, in order to
quantitatively reproduce the disk structure (gaps and rings) with
DALI we need to set different dust scaling factors for the two
gaps: δG1 ∼ 0.1 and δG2 ∼ 0.01.
5. Comparison to hydrodynamical simulations
In this section we investigate the possibility that the gaps were
produced by planets, as has been suggested for other disks, e.g.
HL Tau, TW Hydra, HD 163296, and HD 169142 (ALMA Part-
nership et al. 2015; Andrews et al. 2016; Isella et al. 2016;
Fedele et al. 2017). To do this, we have run 2D simulations of
planet-disk interaction using a version of FARGO-3D (Benítez-
Llambay & Masset 2016) modified to include dust dynamics
(Rosotti et al. 2016). The disk parameters were chosen to match
the best fit model (see Table 3). The simulations were run us-
ing a logarithmic radial grid, extending from 10 au to 300 au at
a resolution of Nr × Nφ = 550 × 1024. Given the relative ex-
pense of the hydrodynamic simulations, we have not conducted
an exhaustive fit to the data, instead we investigate the typical
planetary properties and disk conditions that produce reasonable
gap structures. The key information available for constraining
the planets properties are the gap width, depth and location. The
position of the two gap centers (62 and 103 au) are close enough
to the ratio of radii expected for two planets migrating together
in 2:1 resonance (with semi-major axis ratio 0.63), which forms
the starting point for our investigation.
Starting from the relationship between the width of a gap opened
by a planet and its mass derived by Rosotti et al. (2016), we al-
ready see that the inner planet must be low mass due because the
inner gap is narrow (just 16 au or approximately 2 – 3 pressure
Article number, page 5 of 9
A&A proofs: manuscript no. as209
Fig. 5. Results of hydrodynamical simulations for the inviscid (left) and α = 10−4 (right) case. The black line corresponds to the dust surface
density determined with dali (Sect. 4). The (red) dot-dashed line is the 1 mm grains surface density based on the single planet scenario with 0.2 MJ
planet at 95 au injected after 0.65 Myr (Sect.5). The (blue) dot-dot-dashed and the (green) dashed curves corresponds to the two planets scenario
with an inner planet of mass 0.05 MJ and 0.1 MJ, respectively.
scale-heights). This suggests that the mass of the inner planet is
in the Neptune mass regime (∼ 15 M⊕, i.e. 0.05 MJ). However,
while Rosotti et al. (2016) showed that these planets can produce
observable features, they found gap depths much smaller than
the inner gap in AS 209. Nonetheless, the depth of the gap is sen-
stive to disk viscosity, with planets opening deeper gaps in low
viscosity disks (Crida et al. 2006; Zhu et al. 2013). Thus together
with the gap width, the gap depth places constraints on both the
planet mass and disk viscosity. Furthermore, at extremely low
viscosity, Dong et al. (2017) and Bae et al. (2017) showed that
planets may open multiple dust gaps in inviscid disks, which
raises the interesting possibility that the gaps in AS 209 may
be opened by a single planet.
From the right panel of Fig. 5, we see that even a modest vis-
cosity α = 10−4 is too high to produce deep enough gaps to
explain the structures in AS 209. While a single 0.2 MJ planet at
95 au matches the width of the outer gap, the drop in dust sur-
face density it is far too shallow compared to what is inferred.
Although more massive planets can open deep enough gaps (for
α = 10−4), they produce structures that are too wide and begin
to prevent the inflow of dust entirely. This suggests that the outer
gap is consistent with a 0.2 MJ planet, but requires even lower
viscosity.
The inviscid (α = 0) simulations produce a much better match
to the inferred gap structure (Fig. 5, left panel). Already a single
0.2 MJ planet at 95 au is in remarkable agreement with the gap
structure, producing a deep outer gap along with a second inner
gap at approximately the the right location and with a reason-
able width and depth. The depth of the gaps is not only sensitive
to viscosity, but in the inviscid case it also increases with time,
thus being a much weaker tracer of the planet mass than the po-
sition of the peaks (as noted by Rosotti et al. 2016). However,
Article number, page 6 of 9
the width of the gaps are similar in both the viscous and inviscid
cases. Interestingly, the simulations produce an additional gap
at around 35 au, which coincides with similar, but much smaller
amplitude, feature in the radial intensity profile (Fig. 2). Given
that the depth of this feature is dependent on parameters such as
viscosity, and optical depth effects may further reduce the ampli-
tude of variation in the observed intensity profile, it is possible
that this inner structure is related to the presence of a planet near
100 au.
Since low viscosity is required to reproduce the gap structure in
AS 209, and in this case a single planet may produce both gaps,
it is interesting to consider whether there still could be a second
planet present. Thus we have re-run both the viscous and invis-
cid simulations with an inner planets at 57 au, close to the 2:1
resonance with the outer planet. In both cases the planet mass is
consistent with the estimate from the relationship between gap
width and planet mass: a 0.05 MJ planet produces a negligible
difference to the structure of the gaps and could thus be easily
hidden in the gap. The largest planet mass compatible with the
gap widths is roughly 0.1 MJ, with larger planet masses produc-
ing a gap that is too wide in the inviscid case. While a slightly
more massive planet may be compatible with the inner gap when
α = 10−4, this is hard to reconcile with the need for a lower α in
the outer gap.
6. Conclusions
The ringed dust structure of AS 209 revealed by ALMA is con-
sistent with the presence of a Saturn-like (Mplanet = 0.2MJ =
0.67 MSaturn) planet at r = 95 au. The planetary mass is con-
strained by the width and depth of the dust gap: for the chosen
gas properties, our hydrodynamical simulations indicate that less
D. Fedele et al.: ALMA continuum observations of the protoplanetary disk AS 209
Table 3. Disk model used in the planet disk-interaction simulations
Parameter
Temperature
Gas Surface Density
Grain size
Viscous α
Value
190(R/100 au)−0.55
7.5(R/100 au)−1
1 mm
0, 10−4
massive planets ((cid:46) 0.5 MSaturn) do not produce a gap that is suf-
ficiently wide, while more massive planets ((cid:38) 1.0 MSaturn) pre-
vent the transport of dust inwards entirely, forming a transition
disk (inner hole in large grains). Our radiative transfer calcula-
tions (Sect 4) show that surface density of the mm grains beyond
the planet orbit is enhanced and largely confined in a narrow
ring. This may be the signature of dust pile-up due to the planet-
induced gas pressure maximum beyond the orbit of the planet.
Interestingly, the ALMA C18O image presented in Huang et al.
(2016) shows an extended emission peaking at nearly 130 au.
This extended emission is co-spatial to the outer dust ring. This
is a strong indication that the large scale C18O emission follows
the actual disk surface density in the outer disk and it points to a
gas density drop by a factor of a few inside the outer gap ('G2').
We have also investigated the existence of a second planet in
correspondence of the inner dust gap. We conclude that there
could be a second planet present in the inner gap if it is less than
about 0.1 MJ. Since the two gaps are close to the 2:1 resonance
this raises the possibility that the structure could be caused by a
pair of planets migrating in resonance. The 1 mm surface density
resulting from a pair of 0.05 MJ - 0.2 MJ is in remarkable agree-
ment with the observations. Nonetheless, while both scenarios
require low disk turbulence, the presence of the inner planet is
not needed to explain the observed structures.
The inferred presence of the Saturn-like planet at r ∼ 95 au
raises new questions about planet formation at such large dis-
tance from the star. Gravitational instability can occur on short
timescales and is a viable process for the formation of giant plan-
ets on wide orbits (e.g, Kratter & Lodato 2016). An alternative
scenario is pebble accretion, in which planets grow through the
accretion of cm-sized (or larger) grains that are weakly coupled
to the gas (Lambrechts & Johansen 2012). In this case however
planet formation is challenged by dust migration: according to
dust evolution models such large dust grains are expected to drift
towards the star in much less than a Myr (e.g., Takeuchi et al.
2005; Brauer et al. 2008), which is difficult to reconcile with
disk lifetimes (a few Myr, e.g., Fedele et al. 2010).
In the case of AS 209, the dust inward migration is likely to be
slowed down by the presence of radial dust traps induced by the
presence of a Saturn-like planet. While this can help to reconcile
the presence of large dust with the disk's lifetime, it must hinder
pebble accretion in the inner disk, restricting this mode of planet
formation to the earliest phases of disk evolution, on timescales
(cid:46) 1 Myr.
Acknowledgements. This paper makes use of the following ALMA data:
ADS/JAO.ALMA#2015.1.00486.S. ALMA is a partnership of ESO (represent-
ing its member states), NSF (USA) and NINS (Japan), together with NRC
(Canada), NSC and ASIAA (Taiwan), and KASI (Republic of Korea), in coop-
eration with the Republic of Chile. The Joint ALMA Observatory is operated by
ESO, AUI/NRAO and NAOJ. DF acknowledges support from the Italian Min-
istry of Education, Universities and Research, project SIR (RBSI14ZRHR). MT
has been supported by the DISCSIM project, grant agreement 341137 funded by
the European Research Council under ERC-2013-ADG.
References
ALMA Partnership, Brogan, C. L., Pérez, L. M., et al. 2015, ApJ, 808, L3
Andrews, S. M., Wilner, D. J., Espaillat, C., et al. 2011, ApJ, 732, 42
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond, C. P. 2009,
ApJ, 700, 1502
Andrews, S. M., Wilner, D. J., Zhu, Z., et al. 2016, ApJ, 820, L40
Bae, J., Zhu, Z., & Hartmann, L. 2017, ArXiv e-prints [arXiv:1706.03066]
Benítez-Llambay, P. & Masset, F. S. 2016, ApJS, 223, 11
Birnstiel, T., Klahr, H., & Ercolano, B. 2012, A&A, 539, A148
Brauer, F., Dullemond, C. P., & Henning, T. 2008, A&A, 480, 859
Bruderer, S. 2013, A&A, 559, A46
Bruderer, S., van Dishoeck, E. F., Doty, S. D., & Herczeg, G. J. 2012, A&A, 541,
Crida, A., Morbidelli, A., & Masset, F. 2006, Icarus, 181, 587
D'Alessio, P., Calvet, N., Hartmann, L., Franco-Hernández, R., & Servín, H.
2006, ApJ, 638, 314
Dong, R., Li, S., Chiang, E., & Li, H. 2017, ApJ, 843, 127
Ercolano, B., Rosotti, G. P., Picogna, G., & Testi, L. 2017, MNRAS, 464, L95
Fedele, D., Carney, M., Hogerheijde, M. R., et al. 2017, A&A, 600, A72
Fedele, D., van den Ancker, M. E., Henning, T., Jayawardhana, R., & Oliveira,
J. M. 2010, A&A, 510, A72
Flock, M., Ruge, J. P., Dzyurkevich, N., et al. 2015, A&A, 574, A68
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2013, PASP, 125,
A91
306
Gaia Collaboration, Brown, A. G. A., Vallenari, A., et al. 2016, A&A, 595, A2
Goodman, J. & Weare, J. 2010, Communications in Applied Mathematics and
Computational Science, Vol. 5, No. 1, p. 65-80, 2010, 5, 65
Huang, J., Öberg, K. I., & Andrews, S. M. 2016, ApJ, 823, L18
Huang, J., Öberg, K. I., Qi, C., et al. 2017, ApJ, 835, 231
Isella, A., Guidi, G., Testi, L., et al. 2016, Physical Review Letters, 117, 251101
Kratter, K. & Lodato, G. 2016, ARA&A, 54, 271
Lambrechts, M. & Johansen, A. 2012, A&A, 544, A32
Loomis, R. A., Öberg, K. I., Andrews, S. M., & MacGregor, M. A. 2017, ApJ,
Natta, A., Testi, L., & Randich, S. 2006, A&A, 452, 245
Okuzumi, S., Momose, M., Sirono, S.-i., Kobayashi, H., & Tanaka, H. 2016,
Papaloizou, J. & Lin, D. N. C. 1984, ApJ, 285, 818
Pérez, L. M., Carpenter, J. M., Chandler, C. J., et al. 2012, ApJ, 760, L17
Pinilla, P., Birnstiel, T., Ricci, L., et al. 2012, A&A, 538, A114
Rosotti, G. P., Juhasz, A., Booth, R. A., & Clarke, C. J. 2016, MNRAS, 459,
Takeuchi, T., Clarke, C. J., & Lin, D. N. C. 2005, ApJ, 627, 286
Tazzari, M., Beaujean,
F., & Testi, L.
2017, ArXiv
e-prints
[arXiv:1709.06999]
Tazzari, M., Testi, L., Ercolano, B., et al. 2016, A&A, 588, A53
Zhang, K., Blake, G. A., & Bergin, E. A. 2015, ApJ, 806, L7
Zhu, Z., Stone, J. M., & Rafikov, R. R. 2013, ApJ, 768, 143
840, 23
ApJ, 821, 82
2790
Article number, page 7 of 9
A&A proofs: manuscript no. as209
Fig. A.1. Results of multi-components geometrical fit with MCMC:
comparison of the observed (gray dots) and bestfit model (red line) de-
projected visibilities.
Appendix A: MCMC
Article number, page 8 of 9
D. Fedele et al.: ALMA continuum observations of the protoplanetary disk AS 209
Fig. A.2. 1D and 2D marginalized distributions of the posterior sampling obtained from MCMC. In the 1D marginalized distributions of each
parameter the vertical dashed lines represent the 16th, 50th and 84th percentiles: the median is taken as the bestfit value, and half the interval
between 16th and 84th percentiles as the uncertainty.
Article number, page 9 of 9
I077.580.082.5RcRc0.260.240.22φ1φ12.12.22.3φ2φ261.662.0RG1RG17.68.08.4hwG1hwG10.020.030.04δG1δG10.7750.800δR1δR1102.8103.2103.6RG2RG215.215.616.0hwG2hwG20.020.03δG2δG24.44.8δR2δR2139.6140.0RR2,outRR2,out7.27.6I01.501.752.00δout77.580.082.5Rc0.260.240.22φ12.12.22.3φ261.662.0RG17.68.08.4hwG10.020.030.04δG10.7750.800δR1102.8103.2103.6RG215.215.616.0hwG20.020.03δG24.44.8δR2139.6140.0RR2,out1.501.752.00δoutδout |
1901.01643 | 2 | 1901 | 2019-04-05T02:33:20 | A Hot Saturn Orbiting An Oscillating Late Subgiant Discovered by TESS | [
"astro-ph.EP",
"astro-ph.SR"
] | We present the discovery of TOI-197.01, the first transiting planet identified by the Transiting Exoplanet Survey Satellite (TESS) for which asteroseismology of the host star is possible. TOI-197 (HIP116158) is a bright (V=8.2 mag), spectroscopically classified subgiant which oscillates with an average frequency of about 430 muHz and displays a clear signature of mixed modes. The oscillation amplitude confirms that the redder TESS bandpass compared to Kepler has a small effect on the oscillations, supporting the expected yield of thousands of solar-like oscillators with TESS 2-minute cadence observations. Asteroseismic modeling yields a robust determination of the host star radius (2.943+/-0.064 Rsun), mass (1.212 +/- 0.074 Msun) and age (4.9+/-1.1 Gyr), and demonstrates that it has just started ascending the red-giant branch. Combining asteroseismology with transit modeling and radial-velocity observations, we show that the planet is a "hot Saturn" (9.17+/-0.33 Rearth) with an orbital period of ~14.3 days, irradiance of 343+/-24 Fearth, moderate mass (60.5 +/- 5.7 Mearth) and density (0.431+/-0.062 gcc). The properties of TOI-197.01 show that the host-star metallicity - planet mass correlation found in sub-Saturns (4-8 Rearth) does not extend to larger radii, indicating that planets in the transition between sub-Saturns and Jupiters follow a relatively narrow range of densities. With a density measured to ~15%, TOI-197.01 is one of the best characterized Saturn-sized planets to date, augmenting the small number of known transiting planets around evolved stars and demonstrating the power of TESS to characterize exoplanets and their host stars using asteroseismology. | astro-ph.EP | astro-ph | Draft version April 8, 2019
Typeset using LATEX twocolumn style in AASTeX62
9
1
0
2
r
p
A
5
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
3
4
6
1
0
.
1
0
9
1
:
v
i
X
r
a
A HOT SATURN ORBITING AN OSCILLATING LATE SUBGIANT DISCOVERED BY TESS
Daniel Huber,1 William J. Chaplin,2, 3 Ashley Chontos,1, 4 Hans Kjeldsen,3, 5 Jørgen Christensen-Dalsgaard,3
Timothy R. Bedding,6, 3 Warrick Ball,2, 3 Rafael Brahm,7, 8, 9 Nestor Espinoza,10 Thomas Henning,10
Andr´es Jord´an,8, 9 Paula Sarkis,10 Emil Knudstrup,3 Simon Albrecht,3 Frank Grundahl,3, 5
Mads Fredslund Andersen,3 Pere L. Pall´e,11, 12 Ian Crossfield,13 Benjamin Fulton,14 Andrew W. Howard,15
Howard T. Isaacson,16 Lauren M. Weiss,1 Rasmus Handberg,3 Mikkel N. Lund,3 Aldo M. Serenelli,17, 18
Jakob Rørsted Mosumgaard,3 Amalie Stokholm,3 Allyson Bieryla,19 Lars A. Buchhave,20 David W. Latham,19
Samuel N. Quinn,19 Eric Gaidos,21 Teruyuki Hirano,22, 1 George R. Ricker,13 Roland K. Vanderspek,13
Sara Seager,13, 23, 24 Jon M. Jenkins,25 Joshua N. Winn,26 H. M. Antia,27 Thierry Appourchaux,28 Sarbani Basu,29
Keaton J. Bell,30, 3 Othman Benomar,31 Alfio Bonanno,32 Derek L. Buzasi,33 Tiago L. Campante,34, 35
Z. C¸ elik Orhan,36 Enrico Corsaro,32 Margarida S. Cunha,34 Guy R. Davies,2, 3 Sebastien Deheuvels,37
Samuel K. Grunblatt,1 Amir Hasanzadeh,38 Maria Pia Di Mauro,39 Rafael A. Garc´ıa,40, 41 Patrick Gaulme,30, 3
L´eo Girardi,42 Joyce A. Guzik,43 Marc Hon,44 Chen Jiang,45 Thomas Kallinger,46 Steven D. Kawaler,47
James S. Kuszlewicz,30, 3 Yveline Lebreton,48, 49 Tanda Li,6, 3 Miles Lucas,47 Mia S. Lundkvist,3, 50
Andrew W. Mann,51 St´ephane Mathis,40, 41 Savita Mathur,11, 12 Anwesh Mazumdar,52 Travis S. Metcalfe,53, 54
Andrea Miglio,2, 3 M´ario J. P. F. G. Monteiro,34, 35 Benoit Mosser,48 Anthony Noll,37 Benard Nsamba,34, 35
Jia Mian Joel Ong,29 S. Ortel,36 Filipe Pereira,34, 35 Pritesh Ranadive,52 Clara R´egulo,11, 12
Tha´ıse S. Rodrigues,42 Ian W. Roxburgh,55 Victor Silva Aguirre,3 Barry Smalley,56 Mathew Schofield,2, 3
S´ergio G. Sousa,34 Keivan G. Stassun,57, 58 Dennis Stello,44, 6, 3 Jamie Tayar,1, 59 Timothy R. White,60
Kuldeep Verma,3 Mathieu Vrard,34 M. Yıldız,36 David Baker,61 Michael Bazot,31 Charles Beichmann,62
Christoph Bergmann,63 Lisa Bugnet,40, 41 Bryson Cale,64 Roberto Carlino,65 Scott M. Cartwright,66
Jessie L. Christiansen,62 David R. Ciardi,62 Orlagh Creevey,67 Jason A. Dittmann,19
Jose-Dias Do Nascimento Jr.,19, 68 Vincent Van Eylen,26 Gabor Fur´esz,13 Jonathan Gagn´e,69 Peter Gao,16
Kosmas Gazeas,70 Frank Giddens,71 Oliver J. Hall,2, 3 Saskia Hekker,30, 3 Michael J. Ireland,60
Natasha Latouf,64 Danny LeBrun,64 Alan M. Levine,13 William Matzko,64 Eva Natinsky,61 Emma Page,61
Peter Plavchan,64 Masoud Mansouri-Samani,65 Sean McCauliff,72 Susan E. Mullally,73 Brendan Orenstein,60
Aylin Garcia Soto,23 Martin Paegert,19 Jennifer L. van Saders,1 Chloe Schnaible,61 David R. Soderblom,73
R´obert Szab´o,74, 75 Angelle Tanner,76 C. G. Tinney,63 Johanna Teske,69, 77, 59 Alexandra Thomas,2, 3
Regner Trampedach,53, 3 Duncan Wright,78 Thomas T. Yuan,61 and Farzaneh Zohrabi76
1Institute for Astronomy, University of Hawai'i, 2680 Woodlawn Drive, Honolulu, HI 96822, USA
2School of Physics and Astronomy, University of Birmingham, Birmingham B15 2TT, UK
3Stellar Astrophysics Centre (SAC), Department of Physics and Astronomy, Aarhus University, Ny Munkegade 120, DK-8000 Aarhus C,
Denmark
4NSF Graduate Research Fellow
5Institute of Theoretical Physics and Astronomy, Vilnius University, Sauletekio av. 3, 10257 Vilnius, Lithuania
6Sydney Institute for Astronomy (SIfA), School of Physics, University of Sydney, NSW 2006, Australia
7Center of Astro-Engineering UC, Pontificia Universidad Cat´olica de Chile, Av. Vicuna Mackenna 4860, 7820436 Macul, Santiago, Chile
8Instituto de Astrof´ısica, Facultad de F´ısica, Pontificia Universidad Cat´olica de Chile
9Millennium Institute of Astrophysics, Av. Vicuna Mackenna 4860, 782-0436 Macul, Santiago, Chile
10Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany
11Instituto de Astrof´ısica de Canarias (IAC), 38205 La Laguna, Tenerife, Spain
12Universidad de La Laguna (ULL), Departamento de Astrof´ısica, E-38206 La Laguna, Tenerife, Spain
13Department of Physics, and Kavli Institute for Astrophysics and Space Research, Massachusetts Institute of Technology, 77
Massachusetts Ave., Cambridge, MA 02139, USA
14NASA Exoplanet Science Institute / Caltech-IPAC, Pasadena, CA 91125, USA
15California Institute of Technology, Pasadena, CA 91125, USA
16Department of Astronomy, UC Berkeley, Berkeley, CA 94720, USA
17Institute of Space Sciences (ICE, CSIC) Campus UAB, Carrer de Can Magrans, s/n, E-08193, Barcelona, Spain
18Institut dEstudis Espacials de Catalunya (IEEC), C/Gran Capita, 2-4, E-08034, Barcelona, Spain
19Center for Astrophysics Harvard & Smithsonian, 60 Garden St., Cambridge, MA 02138, USA
20DTU Space, National Space Institute, Technical University of Denmark, Elektrovej 328, DK-2800 Kgs. Lyngby, Denmark
[email protected]
2
Huber et al.
21Department of Earth Sciences, University of Hawaii at M¯anoa, Honolulu, Hawaii 96822, USA
22Department of Earth and Planetary Sciences, Tokyo Institute of Technology, 2-12-1 Ookayama, Meguro-ku, Tokyo 152-8551, Japan
23Department of Earth, Atmospheric, and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge,
MA 02139, USA
24Department of Aeronautics and Astronautics, Massachusetts Institute of Technology, 77 Massachusetts Ave., Cambridge, MA 02139,
25NASA Ames Research Center, Moffett Field, CA, 94035
USA
26Department of Astrophysical Sciences, Princeton University, 4 Ivy Lane, Princeton, NJ 08544, USA
27Tata Institute of Fundamental Research, Mumbai, India
28Univ. Paris-Sud, Institut d'Astrophysique Spatiale, UMR 8617, CNRS, Batiment 121, 91405 Orsay Cedex, France
29Department of Astronomy, Yale University, P.O. Box 208101, New Haven, CT 06520-8101, USA
30Max-Planck-Institut fur Sonnensystemforschung, Justus-von-Liebig-Weg 3, 37077 Gottingen, Germany
31Center for Space Science, New York University Abu Dhabi, UAE
32INAF - Osservatorio Astrofisico di Catania, via S. Sofia 78, 95123, Catania, Italy
33Dept. of Chemistry & Physics, Florida Gulf Coast University, 10501 FGCU Blvd. S., Fort Myers, FL 33965 USA
34Instituto de Astrof´ısica e Ciencias do Espa¸co, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto, Portugal
35Departamento de F´ısica e Astronomia, Faculdade de Ciencias da Universidade do Porto, Rua do Campo Alegre, s/n, PT4169-007
36Department of Astronomy and Space Sciences, Science Faculty, Ege University, 35100, Bornova, Izmir, Turkey
Porto, Portugal
37IRAP, Universit´e de Toulouse, CNRS, CNES, UPS, Toulouse, France
38Department of Physics, University of Zanjan, Zanjan, Iran
39INAF-IAPS, Istituto di Astrofisica e Planetologia Spaziali, Via del Fosso del Cavaliere 100, I-00133 Roma, Italy
40IRFU, CEA, Universit´e Paris-Saclay, F-91191 Gif-sur-Yvette, France
41AIM, CEA, CNRS, Universit´e Paris-Saclay, Universit´e Paris Diderot, Sorbonne Paris Cit´e, F-91191 Gif-sur-Yvette, France
42Osservatorio Astronomico di Padova -- INAF, Vicolo dellOsservatorio 5, I-35122 Padova, Italy
43Los Alamos National Laboratory, XTD-NTA, MS T-082, Los Alamos, NM 87545 USA
44School of Physics, The University of New South Wales, Sydney NSW 2052, Australia
45School of Physics and Astronomy, Sun Yat-Sen University, Guangzhou, 510275, China
46Institute of Astrophysics, University of Vienna, 1180 Vienna, Austria
47Department of Physics and Astronomy, Iowa State University, Ames, IA 50011 USA
48LESIA, CNRS, Universit´e Pierre et Marie Curie, Universit´e Denis, Diderot, Observatoire de Paris, 92195 Meudon cedex, France
49Univ Rennes, CNRS, IPR (Institut de Physique de Rennes) - UMR 6251, F-35000 Rennes, France
50Zentrum fur Astronomie der Universitat Heidelberg, Landessternwarte, Konigstuhl 12, D-69117 Heidelberg, Germany
51Department of Physics and Astronomy, University of North Carolina at Chapel Hill, Chapel Hill, NC 27599, USA
52Homi Bhabha Centre for Science Education, TIFR, V. N. Purav Marg, Mankhurd, Mumbai 400088, India
53Space Science Institute, 4750 Walnut Street, Suite 205, Boulder CO 80301, USA
54Max-Planck-Institut fur Sonnensystemforschung, Justus-von-Liebig-Weg 3, 37077, Gottingen, Germany
55Astronomy Unit, Queen Mary University of London, Mile End Road, London, E1 4NS, UK
56Astrophysics Group, Lennard-Jones Laboratories, Keele University, Staffordshire ST5 5BG, United Kingdom
57Vanderbilt University, Department of Physics & Astronomy, 6301 Stevenson Center Ln., Nashville, TN 37235, USA
58Vanderbilt Initiative in Data-intensive Astrophysics (VIDA), 6301 Stevenson Center Lane, Nashville, TN 37235, USA
59Hubble Fellow
60Research School of Astronomy and Astrophysics, Australian National University, Canberra, ACT 2611, Australia
61Physics Department, Austin College, Sherman, TX 75090, USA
62Caltech/IPAC-NASA Exoplanet Science Institute, Pasadena, CA 91125, USA
63Exoplanetary Science at UNSW, School of Physics, UNSW Sydney, NSW 2052, Australia
64Department of Physics and Astronomy, George Mason University 4400 University Ave, Fairfax, VA 22030
65SGT Inc/NASA Ames Research Center, Moffett Field, CA, 94035
66Proto-Logic Consulting LLC, Washington, DC 20009, USA
67Universit´e Cote d'Azur, Observatoire de la Cote d'Azur, CNRS, Laboratoire Lagrange, France
68Univ. Federal do Rio G. do Norte, UFRN, Dep. de Fsica, CP 1641, 59072-970, Natal, RN, Brazil
69Carnegie Institution of Washington DTM, 5241 Broad Branch Road NW, Washington, DC 20015, USA
70Section of Astrophysics, Astronomy and Mechanics, Faculty of Physics, National and Kapodistrian University of Athens, GR-15784
Zografos, Athens, Greece
71Missouri State University
72LinkedIn work performed at NASA Ames Research Center, Moffett Field, CA, 94035
73Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21212
TOI-197
3
74MTA CSFK, Konkoly Observatory, Budapest, Konkoly Thege Mikl´os ´ut 15-17, H-1121, Hungary
75MTA CSFK Lendulet Near-Field Cosmology Research Group
76Mississippi State University, Department of Physics & Astronomy, Hilbun Hall, Starkville, MS, 39762, USA
77Observatories of the Carnegie Institution for Science, 813 Santa Barbara Street, Pasadena, CA 91101
78University of Southern Queensland, Toowoomba, Qld 4350, Australia
ABSTRACT
We present the discovery of TOI-197.01, the first transiting planet identified by the Transiting
Exoplanet Survey Satellite (TESS ) for which asteroseismology of the host star is possible. TOI-
197 (HIP 116158) is a bright (V = 8.2 mag), spectroscopically classified subgiant which oscillates
with an average frequency of about 430 µHz and displays a clear signature of mixed modes. The
oscillation amplitude confirms that the redder TESS bandpass compared to Kepler has a small effect
on the oscillations, supporting the expected yield of thousands of solar-like oscillators with TESS
2-minute cadence observations. Asteroseismic modeling yields a robust determination of the host
star radius (R(cid:63) = 2.943 ± 0.064R(cid:12)), mass (M(cid:63) = 1.212 ± 0.074M(cid:12)) and age (4.9 ± 1.1 Gyr), and
demonstrates that it has just started ascending the red-giant branch. Combining asteroseismology
with transit modeling and radial-velocity observations, we show that the planet is a "hot Saturn"
(Rp = 9.17 ± 0.33R⊕) with an orbital period of ∼ 14.3 days, irradiance of F = 343 ± 24F⊕, moderate
mass (Mp = 60.5 ± 5.7M⊕) and density (ρp = 0.431 ± 0.062 g cm−3). The properties of TOI-197.01
show that the host-star metallicity -- planet mass correlation found in sub-Saturns (4 − 8R⊕) does
not extend to larger radii, indicating that planets in the transition between sub-Saturns and Jupiters
follow a relatively narrow range of densities. With a density measured to ∼ 15%, TOI-197.01 is one of
the best characterized Saturn-sized planets to date, augmenting the small number of known transiting
planets around evolved stars and demonstrating the power of TESS to characterize exoplanets and
their host stars using asteroseismology.
Keywords: planets and satellites:
individual (TOI-197) -- stars:
fundamental parameters -- tech-
niques: asteroseismology, photometry, spectroscopy -- TESS -- planetary systems
1. INTRODUCTION
Asteroseismology is one of the major success stories
of the space photometry revolution initiated by CoRoT
(Baglin et al. 2006) and Kepler (Borucki et al. 2010).
The detection of oscillations in thousands of stars has
led to breakthroughs such as the discovery of rapidly
rotating cores in subgiants and red giants, as well as
the systematic measurement of stellar masses, radii and
ages (see Chaplin & Miglio 2013, for a review). As-
teroseismology has also become the "gold standard" for
calibrating more indirect methods to determine stellar
parameters such as surface gravity (log g) from spec-
troscopy (Petigura et al. 2017a) and stellar granulation
(Mathur et al. 2011; Bastien et al. 2013; Kallinger et al.
2016; Corsaro et al. 2017; Bugnet et al. 2018; Pande et al.
2018), and age from rotation periods (gyrochronology,
e.g. Garc´ıa et al. 2014; van Saders et al. 2016).
A remarkable synergy that emerged from space-based
photometry is the systematic characterization of exo-
planet host stars using asteroseismology. Following first
asteroseismic studies of exoplanet host stars using ra-
dial velocities (Bouchy et al. 2005; Bazot et al. 2005),
the Hubble Space Telescope (Gilliland et al. 2011) and
CoRoT (Ballot et al. 2011b; Lebreton & Goupil 2014),
Kepler enabled the systematic characterization of exo-
planets with over 100 detections of oscillations in host
stars to date (Huber et al. 2013a; Lundkvist et al. 2016).
In addition to the more precise characterization of exo-
planet radii and masses (Ballard et al. 2014), the synergy
also enabled systematic constraints on stellar spin-orbit
alignments (Chaplin et al. 2014b; Benomar et al. 2014;
Lund et al. 2014; Campante et al. 2016a) and statistical
inferences on orbital eccentricities through constraints
on the mean stellar density (Sliski & Kipping 2014; Van
Eylen & Albrecht 2015; Van Eylen et al. 2019).
The recently launched NASA TESS Mission (Ricker
et al. 2014) is poised to continue the synergy between as-
teroseismology and exoplanet science. Using dedicated
2-minute cadence observations, TESS is expected to de-
tect oscillations in thousands of main-sequence, sub-
giant and early red-giant stars (Schofield et al. 2018),
and simulations predict that at least 100 of these will
host transiting or non-transiting exoplanets (Campante
et al. 2016b). TESS host stars are on average signif-
icantly brighter than typical Kepler hosts, facilitating
ground-based measurements of planet masses with pre-
4
Huber et al.
cisely characterized exoplanet hosts from asteroseismol-
ogy. While some of the first exoplanets discovered with
TESS orbit stars that have evolved off the main se-
quence (Wang et al. 2019; Brahm et al. 2018; Nielsen
et al. 2019), none of them were amenable to asteroseis-
mology using TESS photometry. Here, we present the
characterization of TESS Object of Interest 197 (TOI-
197, HIP 116158) system, the first discovery by TESS
of a transiting exoplanet around a host star in which
oscillations can be measured.
2. OBSERVATIONS
2.1. TESS Photometry
TESS observed TOI-197 in 2-minute cadence during
Sector 2 of Cycle 1 for 27 days. We used the target
pixel files produced by the TESS Science Processing
Operations Center (Jenkins et al. 2016) as part of the
TESS alerts on November 11 20181. We produced a
light curve using the photometry pipeline2 (Handberg
et al., in prep.) maintained by the TESS Asteroseismic
Science Operations Center (TASOC, Lund et al. 2017),
which is based on software originally developed to gen-
erate light curves for data collected by the K2 Mission
(Lund et al. 2015).
Figure 1a shows the raw light curve obtained from
the TASOC pipeline. The coverage is nearly continu-
ous (duty cycle ∼ 93%), with a ∼ 2 day gap separating
the two spacecraft orbits in the observing sector. Two
∼ 0.1 % brightness dips, which triggered the identifica-
tion of TOI-197.01 as a planet candidate, are evident
near the beginning of each TESS orbit (see upward tri-
angles in Figure 1a). The structure with a period of
∼ 2.5 d corresponds to instrumental variations due to the
angular momentum dumping cycle of the spacecraft.
To prepare the raw light curve for an asteroseismic
analysis, the current TASOC pipeline implements a se-
ries of corrections as described by Handberg & Lund
(2014), which includes removal of instrumental arte-
facts and of the transit events using a combination of
filters utilizing the estimated planetary period. Future
TASOC-prepared light curves from full TESS data re-
leases will use information from the ensemble of stars to
remove common instrumental systematics (Lund et al,
in prep.). Alternative light curve corrections using tran-
sit removal and gap interpolation (Garc´ıa et al. 2011;
Pires et al. 2015) yielded consistent results. The cor-
rected TASOC light curve is shown in Figure 1b. Figure
1c shows a power spectrum of this light curve, revealing
1 https://doi.org/10.17909/t9-wx1n-aw08
2 https://tasoc.dk/code/
the clear presence of a granulation background and a
power excess from solar-like oscillations near ∼ 430 µHz,
both characteristic of an evolved star near the base of
the red-giant branch.
2.2. High-Resolution Spectroscopy
We obtained high-resolution spectra of TOI-197 us-
ing several facilities within the TESS Follow-up Obser-
vation Program (TFOP), including HIRES (Vogt et al.
1994) on the 10-m telescope at Keck Observatory (Mau-
nakea, Hawai'i), the Hertzsprung SONG Telescope at
Teide Observatory (Tenerife) (Grundahl et al. 2017),
HARPS (Mayor et al. 2003), FEROS (Kaufer et al.
1999), Coralie (Queloz et al. 2001) and FIDEOS (Vanzi
et al. 2018) on the MPG/ESO 3.6-m, 2.2-m, 1.2-m, and
1-m telescopes at La Silla Observatory (Chile), Veloce
(Gilbert et al. 2018) on the 3.9-m Anglo-Australian Tele-
scope at Siding Spring Observatory (Australia), TRES
(Fur´esz 2008) on the 1.5-m Tillinghast reflector at the
F. L. Whipple Observatory (Mt. Hopkins, Arizona),
and iSHELL (Rayner et al. 2012) on the NASA IRTF
Telescope (Maunakea, Hawaii). All spectra used in this
paper were obtained between Nov 11 and Dec 30 2018
and have a minimum spectral resolution of R ≈ 44000.
FEROS, Coralie, and HARPS data were processed and
analyzed with the CERES package (Brahm et al. 2017a),
which performs the optimal extraction and wavelength
calibration of each spectrum, along with the measure-
ment of precision radial velocities and bisector spans via
the cross-correlation technique. Most instruments have
been previously used to obtain precise radial velocities
to confirm exoplanets, and we refer to the publications
listed above for details on the reduction methods.
To obtain stellar parameters, we analyzed a HIRES
spectrum using Specmatch (Petigura 2015), which has
been extensively applied for the classification of Kepler
exoplanet host stars (Johnson et al. 2017; Petigura et al.
2017a). The resulting parameters were Teff = 5080 ±
70 K, log g = 3.60 ± 0.08 dex, [Fe/H] = −0.08 ± 0.05 dex
and v sin i = 2.8±1.6 km s−1, consistent with an evolved
star as identified from the power spectrum in Figure 1c.
To account for systematic differences between spectro-
scopic methods (Torres et al. 2012) we added 59 K in Teff
and 0.062 dex in [Fe/H] in quadrature to the formal un-
certainties, yielding final values of Teff = 5080±90 K and
[Fe/H] = −0.08 ± 0.08 dex. Independent spectroscopic
analyses yielded consistent results, including an analy-
sis of a HIRES spectrum using ARES+MOOG (Sousa
2014; Sousa et al. 2018), FEROS spectra using ZASPE
(Brahm et al. 2017b), TRES spectra using SPC (Buch-
have et al. 2012) and iSHELL spectra using BT-Settl
models (Allard et al. 2012).
TOI-197
5
Figure 1. Panel (a): Raw TESS 2-minute cadence light curve of TOI-197 produced by the TESS Asteroseismic Science
Operations Center (TASOC). The red line is the light curve smoothed with a 10-minute boxcar filter (shown for illustration
purposes only). Upward triangles mark the two transit events. Panel (b): Light curve after applying corrections by the TASOC
pipeline. Panel (c): Power spectrum of panel (b), showing a granulation background and power excess due to oscillations near
∼ 430 µHz. The solid red line is a global fit, consisting of granulation plus white noise and a Gaussian describing the power
excess due to oscillations. Dashed red lines show the two granulation components and the white noise level, respectively.
2.3. Broadband Photometry & Gaia Parallax
We fitted the spectral energy distribution (SED) of
TOI-197 using broadband photometry following the
method described by Stassun & Torres (2016). We used
NUV photometry from GALEX, BT VT from Tycho-2
(Høg et al. 2000), BV gri from APASS, JHKS from
2MASS (Skrutskie et al. 2006), W1 -- W4 from WISE
(Wright et al. 2010), and the G magnitude from Gaia
(Evans et al. 2018). The data were fit using Kurucz
atmosphere models, with Teff ,
[Fe/H] and extinc-
tion (AV ) as free parameters. We restricted AV to
the maximum line-of-sight value from the dust maps
of Schlegel et al. (1998). The resulting fit yielded
Teff = 5090 ± 85 K,
[Fe/H] = −0.3 ± 0.3 dex, and
AV = 0.09 ± 0.02 mag with reduced χ2 of 1.9,
in
good agreement with spectroscopy. Integrating the (de-
reddened) model SED gives the bolometric flux at Earth
of Fbol = 1.88±0.04×10−8 erg s cm−2. An independent
SED fit using 2MASS, APASS9, USNO-B1 and WISE
photometry and Kurucz models yielded excellent agree-
ment, with Fbol = 1.83 ± 0.09 × 10−8 erg s cm−2 and
Teff = 5150 ± 130 K. Additional independent analyses
135513601365137013751380BJD-2457000 (days)200002000Flux (ppm)(a)135513601365137013751380BJD-2457000 (days)20001000010002000Flux (ppm)(b)102103Frequency (Hz)101102103Power Density (ppm2 Hz1)(c)6
Huber et al.
using the method by Mann et al. (2016) and PARAM
(Rodrigues et al. 2014, 2017) yielded bolometric fluxes
and extinction values that are consistent within 1 σ with
the values quoted above.
Combining the bolometric flux with the Gaia DR2
distance allows us to derive a nearly model-independent
luminosity, which is a valuable constraint for asteroseis-
mic modeling (see Section 3.3). Using a Gaia parallax of
10.518 ± 0.080 mas (adjusted for the 0.082 ± 0.033 mas
zero-point offset for nearby stars reported by Stassun
& Torres 2018) with the two methods described above
yielded L(cid:63) = 5.30 ± 0.14L(cid:12) (using Fbol = 1.88 ±
0.04 × 10−8 erg s cm−2) and L(cid:63) = 5.13 ± 0.13L(cid:12) (using
Fbol = 1.83 ± 0.09 × 10−8 erg s cm−2). We also derived
a luminosity using isoclassify (Huber et al. 2017)3,
adopting 2MASS K-band photometry, bolometric cor-
rections from MIST isochrones (Choi et al. 2016) and
the composite reddening map mwdust (Bovy et al. 2016),
yielding L(cid:63) = 5.03 ± 0.13L(cid:12). Our adopted luminosity
was the mean of these methods with an uncertainty cal-
culated by adding the mean uncertainty and scatter over
all methods in quadrature, yielding L(cid:63) = 5.15± 0.17L(cid:12).
2.4. High-Resolution Imaging
TOI-197 was observed with the NIRC2 camera and
Altair adaptive optics system on Keck-2 (Wizinowich
et al. 2000) on UT 25 November 2018. Conditions were
clear but seeing was poor (0.8 -- 2"). We used the science
target as the natural guide star and images were ob-
tained through a K-continuum plus KP501.5 filter using
the narrow camera (10 mas pixel scale). We obtained
eight images (four each at two dither positions), each
consisting of 50 co-adds of 0.2 sec each, with correlated
double-sampling mode and four reads. Frames were co-
added and we subtracted an average dark image, con-
structed from a set of darks with the same integration
time and sampling mode. Flat-fielding was performed
using a dome flat obtained in the K(cid:48) filter. "Hot" pixels
were identified in the dark image and corrected by me-
dian filtering with a 5 × 5 box centered on each affected
pixel in the science image. Only a single star appears
in the images. We performed tests in which "clones" of
the stellar image reduced by a specified contrast ratio
were added to the original image. These show that we
would have been able to detect companions as faint as
∆K = 5.8 mag within 0.4" of TOI-197, 3.8 mag within
0.2", and 1.8 mag within 0.1".
Additional NIRC2 observations were obtained in the
narrow-band Br−γ filter (λo = 2.1686; ∆λ = 0.0326µm)
on UT 22 November 2018. A standard 3-point dither
pattern with a step size of 3(cid:48)(cid:48) was repeated twice with
each dither offset from the previous dither by 0.5(cid:48)(cid:48). An
integration time of 0.25 seconds was used with one coadd
per frame for a total of 2.25 seconds on target, and the
camera was used in the narrow-angle mode. No addi-
tional stellar companions were detected to within a res-
olution of ∼ 0.05(cid:48)(cid:48) FWHM. The sensitivities of the final
combined AO image were determined following Ciardi
et al. (2015) and Furlan et al. (2017), with detection lim-
its as faint as ∆Br − γ = 7.4 mag within 0.4", 6.1 mag
within 0.2", and 3.2 mag within 0.1".
The results from NIRC2 are consistent with Speckle
observations using HRCam (Tokovinin et al. 2010) on
the 4.1 m SOAR telescope4. Since the companion is un-
likely to be bluer than TOI-197, these constraints ex-
clude any significant dilution (both for oscillation am-
plitudes and the depth of transit events).
3. ASTEROSEISMOLOGY
3.1. Global Oscillation Parameters
To extract oscillation parameters characterizing the
average properties of the power spectrum, we used sev-
eral automated analysis methods (e.g. Huber et al. 2009;
Mathur et al. 2010; Mosser et al. 2012a; Benomar et al.
2012; Kallinger et al. 2012; Corsaro & De Ridder 2014;
Lundkvist 2015; Stello et al. 2017; Campante 2018; Bell
et al. 2019), many of which have been extensively tested
on Kepler data (e.g. Hekker et al. 2011; Verner et al.
2011). In most of these analyses, the power contribu-
tions due to granulation noise and stellar activity were
modeled by a combination of power laws and a flat con-
tribution due to shot noise, and then corrected by di-
viding the power spectrum by the background model.
The individual contributions and background model us-
ing the method by Huber et al. (2009) are shown as
dashed and solid red lines in Fig. 1c, and a close-up of
the power excess is shown in Fig. 2a.
Next, the frequency of maximum power (νmax) was
measured either by heavily smoothing the power spec-
trum or by fitting a Gaussian function to the power
excess. Our analysis yielded νmax = 430 ± 18 µHz,
with uncertainties calculated from the scatter between
all fitting techniques. Finally, the mean oscillation
amplitude per radial mode was determined by taking
the peak of the smoothed, background-corrected os-
cillation envelope and correcting for the contribution
of non-radial modes (Kjeldsen et al. 2008b), yielding
A = 18.7 ± 3.5 ppm. We caution that the νmax and
3 https://github.com/danxhuber/isoclassify
4
441462736
https://exofop.ipac.caltech.edu/tess/target.php?id=
TOI-197
7
Figure 2. Panel (a): Power spectrum of TOI-197 cen-
tered on the frequency region showing oscillations. Vertical
dashed lines mark identified individual frequencies. Panel
(b): Greyscale ´echelle diagram5 of the background-corrected
and smoothed power spectrum in panel (a). Identified indi-
vidual mode frequencies are marked with blue circles (l = 0,
radial modes), green squares (l = 2, quadrupole modes) and
red diamonds (l = 1, dipole modes). Note that the diagram
is replicated for clarity (Bedding 2012).
amplitude estimates could be significantly biased by the
stochastic nature of the oscillations. The modes are not
well resolved, as demonstrated by the non-Gaussian ap-
pearance of the power spectrum and the particularly
strong peak at 420 µHz.
Global seismic parameters such as νmax and ampli-
tude follow well-known scaling relations (Huber et al.
2011; Mosser et al. 2012b; Corsaro et al. 2013), which
allow us to test whether the detected oscillations are
consistent with expectations. Figure 3 compares our
measured νmax and amplitude with results for ∼1500
stars observed by Kepler (Huber et al. 2011). We ob-
serve excellent agreement, confirming that the detected
5 ´Echelle diagrams are constructed by dividing a power spec-
trum into equal segments with length ∆ν and stacking one above
the other, so that modes with a given spherical degree align verti-
cally in ridges (Grec et al. 1983). Departures from regularity arise
from sound speed discontinuities and from mixed modes, and thus
probe the interior structure of a star.
Figure 3. Amplitude per radial mode versus frequency
of maximum power for a sample of ∼ 1500 stars spanning
from the main-sequence to the red giant branch observed by
Kepler (Huber et al. 2011). The red star shows the measured
position of TOI-197. The uncertainties are approximately
equal to the symbol size.
signal is consistent with solar-like oscillations. We note
that the oscillations in the TESS bandpass are expected
to be ∼ 15 % smaller than in the bluer Kepler band-
pass, which is well within the spread of amplitudes at a
given νmax observed in the Kepler sample. The result
confirms that the redder bandpass of TESS only has a
small effect on the oscillation amplitude, supporting the
expected rich yield of solar-like oscillators with TESS
2-minute cadence observations (Schofield et al. 2018).
3.2. Individual Mode Frequencies
The power spectrum in Fig. 2a shows several clear
peaks corresponding to individual oscillation modes.
Given that TESS instrument artifacts are not yet well
understood, we restricted our analysis to the frequency
range 400 -- 500 µHz where we observe peaks well above
the background level.
To extract these individual mode frequencies, we
independent methods ranging from tra-
used several
i.e., pre-whitening
ditional iterative sine-wave fitting,
(e.g. Lenz & Breger 2005; Kjeldsen et al. 2005; Bed-
ding et al. 2007), to fitting of Lorentzian mode profiles
(e.g. Handberg & Campante 2011; Appourchaux et al.
2012; Mosser et al. 2012b; Corsaro & De Ridder 2014;
Corsaro et al. 2015; Vrard et al. 2015; Davies & Miglio
2016; Roxburgh 2017; Handberg et al. 2017; Kallinger
et al. 2018), including publicly available code such as
DIAMONDS6. We required at least two independent meth-
6https://github.com/EnricoCorsaro/DIAMONDS
380400420440460480500520Frequency (Hz)05001000150020002500Power Density (ppm2Hz1)(a)0102001020Frequency mod 28.94 Hz380400420440460480500520Frequency (Hz)(b)101102103Frequency of Maximum Power(Hz)101102103Amplitude (ppm)TOI-1978
Huber et al.
Table 1. Extracted oscillation frequencies and mode iden-
tifications for TOI-197.
f (µHz) σf (µHz)
413.12
420.06
429.26
436.77
445.85
448.89
460.16
463.81
477.08
478.07
0.29
0.11
0.14
0.24
0.21
0.21
0.33
0.43
0.31
0.35
l
1
0
1
1
2
0
1
1
1
0
Note: The large frequency separation derived from radial
modes is ∆ν = 28.94 ± 0.15µHz. Note that the l = 1 modes
at ∼ 460 and ∼ 463 µHz are listed for completeness, but it is
unlikely that both of them are genuine (see text).
ods to return the same frequency within uncertainties
and that the posterior probability of each peak being
a mode was ≥ 90 % (Basu & Chaplin 2017). A com-
parison of the frequencies returned by different fitters
showed very good agreement, at a level smaller than the
uncertainties for all the reported modes. For the final
list of frequencies we adopted values from one fitter who
applied pre-whitening (HK), with uncertainties derived
from Monte Carlo simulations of the data, as listed in
Table 1.
To measure the large frequency separation ∆ν, we per-
formed a linear fit to all identified radial modes, yield-
ing ∆ν = 28.94 ± 0.15µHz. Figure 2b shows a greyscale
´echelle diagram5 using this ∆ν measurement, including
the extracted mode frequencies. The l = 1 modes are
strongly affected by mode bumping, as expected for the
mixed mode coupling factors for evolved stars in this
evolutionary stage. The offset of the l = 0 ridge is
∼ 1.5, consistent with the expected value from Kepler
measurements for stars with similar ∆ν and Teff (White
et al. 2011).
3.3. Frequency Modeling
We used a number of
independent approaches to
model the observed oscillation frequencies,
including
different stellar evolution codes (ASTEC, Cesam2K,
GARSTEC, Iben, MESA, and YREC, Christensen-
Dalsgaard 2008; Morel & Lebreton 2008; Scuflaire
et al. 2008; Weiss et al. 2008; Iben 1965; Paxton et al.
2011, 2013, 2015; Choi et al. 2016; Demarque et al.
2008), oscillation codes (ADIPLS, GYRE and Pesnell,
Christensen-Dalsgaard 2008; Townsend & Teitler 2013;
Pesnell 1990) and modeling methods (including AMP,
´Echelle diagram showing observed oscillation
Figure 4.
frequencies (filled grey symbols) and a representative best-
fitting model (open colored symbols) using GARSTEC,
ADIPLS and BeSSP (Serenelli et al. 2017). Model symbol
sizes for non-radial modes are scaled using the mode inertia
(a proxy for mode amplitude) as described in Cunha et al.
(2015). Thick model symbols correspond to modes that were
matched to observations. Uncertainties on the observed fre-
quencies are than smaller or comparable to the symbol sizes.
Note that the l = 1 mode at 460 µHz has been omitted from
this plot (see text).
ASTFIT, BeSSP, BASTA, PARAM, Creevey et al. 2017;
Silva Aguirre et al. 2015; Serenelli et al. 2017; Rodrigues
et al. 2014, 2017; Deheuvels & Michel 2011; Yıldız et al.
2016; Ong & Basu 2019; Tayar & Pinsonneault 2018; Le-
breton & Goupil 2014; Ball & Gizon 2017; Mosumgaard
et al. 2018). Most of the adopted methods applied cor-
rections for the surface effect (Kjeldsen et al. 2008a; Ball
& Gizon 2017). Model inputs included the spectroscopic
temperature and metallicity, individual frequencies, ∆ν,
and the luminosity (Section 2.3). To investigate the ef-
fects of different input parameters, modelers were asked
to provide solutions using both individual frequencies
and only using ∆ν, with and without taking into ac-
count the luminosity constraint. The constraint on νmax
was not used in the modeling since it may be affected
by finite mode lifetimes (see Section 3.1).
Overall, the modeling efforts yielded consistent re-
sults, and most modeling codes were able to provide ad-
equate fits to the observed oscillation frequencies. The
0510152025Frequency mod 28.9 Hz300350400450500550Frequency (Hz)l=0l=1l=2TOI-197
9
Table 2. Host Star Parameters
Table 3. High-precision Radial Velocities for TOI-197
Basic Properties
Time (BJD)
RV (m/s) σRV (m/s)
Instrument
Hipparcos ID
TIC ID
V Magnitude
TESS Magnitude
K Magnitude
SED & Gaia Parallax
Parallax, π (mas)
Luminosity, L (L(cid:12))
116158
441462736
8.15
7.30
6.04
10.518 ± 0.080
5.15 ± 0.17
Spectroscopy
Effective Temperature, Teff (K)
Metallicity, [Fe/H] (dex)
Projected rotation speed, v sin i (km s−1)
5080 ± 90
−0.08 ± 0.08
2.8 ± 1.6
Asteroseismology
Stellar Mass, M(cid:63) (M(cid:12))
Stellar Radius, R(cid:63) (R(cid:12))
Stellar Density, ρ(cid:63) (gcc)
Surface gravity, log g (cgs)
Age, t (Gyr)
1.212 ± 0.074
2.943 ± 0.064
4.9 ± 1.1
0.06702 ± 0.00067
3.584 ± 0.010
Notes: The TESS magnitude is adopted from the TESS
Input Catalog (Stassun et al. 2018).
modeling confirmed that only one of the two closely-
spaced mixed modes near ∼ 460 µHz is likely real, but
we have retained both frequencies in Table 1 for consis-
tency. An ´echelle diagram with observed frequencies and
a representative best-fitting model is shown in Figure 4.
Independent analyses confirmed a bimodality splitting
into lower-mass, older models (∼ 1.15M(cid:12), ∼ 6 Gyr) and
higher-mass, younger models (∼ 1.3M(cid:12), ∼ 4 Gyr). Sur-
face rotation would provide an independent mass diag-
nostic (e.g. van Saders & Pinsonneault 2013), but the
insufficiently constrained v sin i and the unknown stellar
inclination mean that we cannot decisively break this
degeneracy. Combining an independent constraint of
log g = 3.603 ± 0.026 dex from an autocorrelation anal-
ysis of the light curve (Kallinger et al. 2016) with a
radius from L and Teff favors a higher-mass solution
(M(cid:63) = 1.27 ± 0.13M(cid:12)), but may be prone to small sys-
tematics in the νmax scaling relation (which was used
for the calibration). To make use of the most obser-
vational constraints available, we used the set of nine
modeling solutions which used Teff , [Fe/H], frequencies
and the luminosity as input parameters. From this set
of solutions, we adopted the self-consistent set of stellar
2458426.334584
2458426.503655
2458427.575230
2458428.547576
. . .
2458443.535340
2458443.541210
2458443.714865
2458443.825283
. . .
2458482.562290
2458483.541710
2458483.553240
2458483.564690
4.258
6.328
-12.667
17.328
. . .
-14.667
-3.067
-6.815
-4.375
. . .
19.433
16.133
19.233
16.233
11.260
11.270
3.000
18.540
. . .
3.600
3.800
0.780
0.720
. . .
2.000
2.000
2.000
2.000
SONG
SONG
FEROS
SONG
. . .
CORALIE
CORALIE
HIRES
HIRES
. . .
HARPS
HARPS
HARPS
HARPS
Notes: Error bars do not include contributions from stellar
jitter and measurements have not been corrected for
zeropoint offsets. This table is available in its entirety in a
machine-readable form in the online journal.
parameters with the mass closest to the median mass
over all results. A more detailed study of the individual
modeling results will be presented in a follow-up paper
(Li et al., in prep).
For ease of propagating stellar parameters to exo-
planet modeling (see next section), uncertainties were
calculated by adding the median uncertainty for a given
stellar parameter in quadrature to the standard devi-
ation of the parameter for all methods. This method
has been commonly adopted for Kepler (e.g. Chaplin
et al. 2014a) and captures both random and system-
atic errors estimated from the spread among different
methods. For completeness, the individual random and
systematic error estimates are R(cid:63) = 2.943±0.041(ran)±
0.049(sys) R(cid:12), M(cid:63) = 1.212±0.052(ran)±0.055(sys) M(cid:12),
ρ(cid:63) = 0.06702 ± 0.00019(ran) ± 0.00047(sys) gcc, and
t = 4.9 ± 0.6(ran) ± 0.9(sys) Gyr. This demonstrates
that systematic errors constitute a significant fraction of
the error budget for all stellar properties (in particular
stellar age), and emphasize the need for using multiple
model grids to derive realistic uncertainties for stars and
exoplanets. The final estimates of stellar parameters are
summarized in Table 2, constraining the radius, mass,
density and age of TOI-197 to ∼ 2 %, ∼ 6 %, ∼ 1 % and
∼ 22 %.
4. PLANET CHARACTERIZATION
To fit the transits observed in the TESS data we used
the PDC-MAP light curve provided by the TESS Sci-
ence Processing and Operations Center (SPOC), which
10
Huber et al.
Figure 5. Radial velocity timeseries (panel a) and residuals after subtracting the best-fitting model (panel b) for TOI-197.
Datapoints are corrected for zeropoint offsets of individual instruments, and error bars include contributions from stellar jitter.
has been optimized to remove instrumental variability
and preserve transits (Smith et al. 2012; Stumpe et al.
2014). To optimize computation time we discarded all
data more than 2.5 days before and after each of the
two observed transits. We have repeated the fit and data
preparation procedure using the TASOC light curve and
found consistent results.
A total of 107 radial velocity measurements from five
different instruments (see Section 2.2 and Table 3) were
used to constrain the mass of the planet. No spec-
troscopic observations were taken during transits, and
hence the measurements are unaffected by the Rossiter-
McLaughlin effect (∼ 2.3 m s−1 based on the measured
v sin i and Rp/R(cid:63)). To remove variations from stellar
oscillations, we calculated weighted nightly means for
all instruments which obtained multiple observations
per night. We performed a joint transit and radial-
velocity fit using a Markov Chain Monte Carlo algorithm
based on the exoplanet modeling code ktransit (Bar-
clay 2018), as described in Chontos et al. (2019). We
placed a strong Gaussian prior on the mean stellar den-
sity using the value derived from asteroseismology (Ta-
ble 2) and weak priors on the linear and quadratic limb
darkening coefficients, derived from the closest I-band
grid points in Claret & Bloemen (2011), with a width of
0.6 for both coefficients. We also adopted a prior for the
radial-velocity jitter from granulation and oscillations
of 2.5 ± 1.5 m s−1, following Yu et al. (2018) (see also
Tayar et al. 2018), and a 1/e prior on the eccentricity
to account for the linear bias introduced by sampling in
e cos ω and e sin ω (Eastman et al. 2013). Independent
zeropoint offsets and stellar jitter values for each of the
five instruments that provided radial velocities.
Inde-
pendent joint fits using EXOFASTv2 (Eastman et al.
2013) yielded consistent results.
Figures 5 and 6 show the radial velocity timeseries,
phase-folded transit and RV data, and the corresponding
best-fitting model. Table 4 lists the summary statistics
for all planet and model parameters. The system is well
described by a planet in a 14.3 day orbit, which is nearly
equal in size but ∼ 35% less massive than Saturn (Rp =
0.836 ± 0.031 RJ, Mp = 0.190 ± 0.018 MJ), with tenta-
tive evidence for a mild eccentricity (e = 0.11 ± 0.03).
The long transit duration (∼ 0.5 days) is consistent with
a non-grazing (b ≈ 0.7) transit given the asteroseismic
mean stellar density, providing further confirmation for
a gas-giant planet orbiting an evolved star. The radial
velocity data do not show evidence for any other short-
period companions. Continued monitoring past the ∼4
orbital periods covered here will further reveal details
about the orbital architecture of this system.
5. DISCUSSION
TOI-197.01 joins an enigmatic but growing class of
transiting planets orbiting stars that have significantly
evolved off the main sequence. Figure 7 compares
the position of TOI-197 within the expected popula-
tion of solar-like oscillators to be detected with TESS
(panel a) and within the known population of exoplanet
host stars. Evolutionary states in Figure 7b have been
assigned using solar-metallicity PARSEC evolutionary
tracks (Bressan et al. 2012) as described in Berger et al.
143014401450146014701480BJD - 24570001680816Residuals(b)201001020Radial velocity (m s1)(a)FEROSCORALIESONGHIRESHARPSTOI-197
11
Table 4. Planet Parameters
Parameter Best-fit Median
84%
16%
Model Parameters
γHIRES
γSONG
γFEROS
γCORALIE
γHARPS
σHIRES
σSONG
σFEROS
σCORALIE
σHARPS
z (ppm)
P (days)
T0 (BTJD)
b
Rp/R(cid:63)
e cos ω
e sin ω
K (m/s)
ρ(cid:63)(ρ(cid:12))
u1
u2
e
ω
a (AU)
a/R(cid:63)
i (o)
Rp(R⊕)
Rp( RJ)
Mp(M⊕)
Mp( MJ)
ρp (gcc)
4.8
1.1
-15.4
-5.4
8.1
2.71
2.06
3.49
1.88
2.41
199.4
+1.6
+1.5
+1.2
+1.2
+1.5
+0.85
+0.91
+0.75
+0.75
+0.75
+10.6
−1.6
5.4
−1.5
0.2
−1.2
-15.7
−1.2
-5.0
−1.5
8.8
−0.80
2.68
−0.89
2.11
−0.71
3.47
−0.64
2.50
−0.63
2.69
−10.7
199.1
14.2767 +0.0037 −0.0037
14.2762
1357.0135 1357.0149 +0.0025 −0.0026
−0.049
0.728
0.744
0.02854 +0.00084 −0.00071
−0.061
-0.028
−0.030
-0.096
−1.2
14.1
0.06702 +0.00052 −0.00052
0.35
0.44
−0.24
−0.44
0.02846
-0.054
-0.099
14.6
0.06674
+0.36
+0.30
+0.040
+0.063
+0.029
+1.2
0.12
0.71
Derived Properties
0.113
-118.7
0.1233
9.00
85.67
9.16
0.835
63.4
0.200
0.455
0.115
-106.0
0.1228
8.97
85.75
9.17
0.836
60.5
0.190
0.431
−0.030
+0.034
−31.1
+34.7
+0.0025 −0.0026
−0.27
+0.27
−0.35
+0.36
−0.31
+0.34
−0.028
+0.031
−5.7
−0.018
−0.060
+0.018
+0.064
+5.7
Notes: Parameters denote velocity zeropoints γ, radial
velocity jitter σ, photometric zero point z, orbital period P ,
time of transit T0, impact parameter b, star-to-planet
radius ratio Rp/R(cid:63), eccentricity e, argument of periastron
ω, radial velocity semi-amplitude K, mean stellar density
ρ(cid:63), linear and quadratic limb darkening coefficients u1 and
u2, semi-major axis a, orbital inclination i, as well as planet
radius (Rp), mass (Mp) and density (ρp).
Figure 6. TESS light curve (panel a) and radial-velocity
measurements (panel b) folded with the best fitting orbital
period. Grey points in panel a show the original sampling,
and black points are binned means over 10 minutes. Red lines
in both pabels show the best-fitting model from the joint fit
using stellar parameters, transit and radial velocities. Grey
lines show random draws from the joint MCMC model. Error
bars in panel b include contributions from stellar jitter.
(2018)6 . TOI-197 sits at the boundary between sub-
giants and red giants, with the measured ∆ν value indi-
cating that the star has just started its ascent on the
red-giant branch (Mosser et al. 2014). TOI-197 is a
typical target for which we expect to detect solar-like
oscillations with TESS, predominantly due to the in-
creased oscillation amplitude, which are well known to
scale with luminosity (Kjeldsen & Bedding 1995). On
the contrary, TOI-197 is rare among known exoplanet
hosts: while radial velocity searches have uncovered a
large number of planets orbiting red giants on long or-
bital periods (e.g. Wittenmyer et al. 2011), less than
15 transiting planets are known around red-giant stars
(as defined in Figure 7b). TOI-197 is the sixth example
of a transiting planet orbiting a late subgiant / early
red giant with detected oscillations, following Kepler-91
6see also https://github.com/danxhuber/evolstate
12
Huber et al.
Figure 7. Stellar radius versus effective temperature for the expected TESS Cycle 1 yield of solar-like oscillators (panel a,
Schofield et al. 2018) and for all stars with confirmed transiting planets (panel b). The blue dashed line in panel a marks the
approximate limit below which 2-minute cadence data is required to sample the oscillations. Symbols in panel b are color-coded
according to the evolutionary state of the star using solar-metallicity PARSEC evolutionary tracks. TOI-197 falls on the border
between subgiants and red-giants, and is highlighted with an orange/red/blue star symbol. TOI-197 is a typical target for which
we expect to detect solar-like oscillations with TESS, but occupies a rare parameter space for an exoplanet host.
(Barclay et al. 2013), Kepler-56 (Steffen et al. 2012; Hu-
ber et al. 2013b), Kepler-432 (Quinn et al. 2015), K2-97
(Grunblatt et al. 2016) and K2-132 (Grunblatt et al.
2017; Jones et al. 2018).
Transiting planets orbiting evolved stars are excellent
systems to advance our understanding of the effects of
stellar evolution on the structure and evolution of plan-
ets (see e.g. Veras 2016, for a review). For example,
such systems provide the possibility to test the effects
of stellar mass, evolution and binarity on planet occur-
rence (e.g. Johnson et al. 2010; Schlaufman & Winn
2013; Stephan et al. 2018), which are still poorly un-
derstood. Furthermore, the increased irradiance on the
planet caused by the evolution of the host star has been
proposed as a means to distinguish between proposed
mechanisms to explain the inflation of gas-giant plan-
ets beyond the limits expected from gravitational con-
traction and cooling (Hubbard et al. 2002; Lopez &
Fortney 2016). Recent discoveries by the K2 mission
have indeed yielded evidence that planets orbiting low-
luminosity RGB stars are consistent with being inflated
by the evolution of the host star (Grunblatt et al. 2016,
2017), favoring scenarios in which the energy from the
star is deposited into the deep planetary interior (Bo-
denheimer et al. 2001).
Based on its radius and orbital period, TOI-197 would
nominally be classified as a warm Saturn, sitting be-
tween the well-known population of hot Jupiters and
the ubiquitous population of sub-Neptunes uncovered
by Kepler (Figure 8a). Taking into account the evolu-
tionary state of the host star, however, TOI-197 falls at
the beginning of the "inflation sequence" in the radius-
incident flux diagram (Figure 8b), with planet radius
strongly increasing with stellar incident flux (Kov´acs
et al. 2010; Demory & Seager 2011; Miller & Fortney
2011; Thorngren & Fortney 2018). Since TOI-197.01
is currently not anomalously large compared to the ob-
served trend and scatter for similar planets (Figure 8b)
and low-mass planets are expected to be more suscepti-
ble to planet reinflation (Lopez & Fortney 2016), TOI-
197 may be a progenitor of a class of re-inflated gas-giant
planets orbiting RGB stars.
If confirmed, the mild eccentricity of TOI-197.01
would be consistent with predictions of a population
of planets around evolved stars for which orbital decay
occurs faster than tidal circularization (Villaver et al.
2014; Grunblatt et al. 2018). Moreover, combining the
asteroseismic age of the system with the possible non-
zero eccentricity would allow constraints on the tidal
dissipation in the planet, which drives the circulariza-
tion of the orbit. Using the formalism by Mardling
(2011) (see also Gizon et al. 2013; Davies et al. 2016;
Ceillier et al. 2016), the current constraints would imply
a minimum value of the planetary tidal quality factor
Qp;min ≈ 3.2 × 104, below which the system would have
been already circularized in ∼ 5 Gyr. Compared to the
4000500060007000Effective Temperature (K)100101Radius (Solar)(a)4000500060007000Effective Temperature (K)100101Radius (Solar)(b)DwarfsSubgiantsGiantsTOI-197
13
Figure 8. Planet radius versus orbital period (panel a) and incident flux (panel b) for confirmed exoplanets. Symbols are
color-coded according to the evolutionary state of the host star (see Figure 7). TOI-197 b is highlighted in both panels with an
orange/red/blue star symbol.
value measured in Saturn (Q ≈ 1800, Lainey et al.
2017), this would demonstrate the broad diversity of
dissipation observed in giant planets. Since tidal dissi-
pation mechanisms vary strongly with internal structure
(see e.g. Guenel et al. 2014; Ogilvie 2014; Andr´e et al.
2017), this may also contribute to understanding the
internal composition of such planets. We caution, how-
ever, that further RV measurements will be needed to
confirm a possible non-zero eccentricity for TOI-197.01.
The precise characterization of planets orbiting
evolved, oscillating stars also provides valuable insights
into the diversity of compositions of planets through
their mean densities. TOI-197.01 falls in the transition
region between Neptune and sub-Saturn sized planets
for which radii increase as RP ≈ M 0.6
P , and Jovian
planets for which radius is nearly constant with mass
(Weiss et al. 2013; Chen & Kipping 2017, Figure 9).
Recent studies of a population of sub-Saturns in the
range ∼ 4 -- 8 R⊕ also found a wide variety of masses,
approximately 6 -- 60 M⊕, regardless of size (Petigura
et al. 2017b; Van Eylen et al. 2018). Furthermore,
masses of sub-Saturns correlate strongly with host star
metallicity, suggesting that metal-rich disks form more
massive planet cores. TOI-197.01 demonstrates that
this trend does not appear to extend to planets with
sizes > 8 R⊕, given its mass of ∼ 60 M⊕ and a roughly
sub-solar metallicity host star ([Fe/H] ≈ −0.08 dex).
This suggests that Saturn-sized planets may follow a
relatively narrow range of densities, a possible signature
of the transition in the interior structure (such as the
increased importance of electron degeneracy pressure,
Zapolsky & Salpeter 1969) leading to different mass-
Figure 9. Mass-radius diagram for confirmed planets with
densities measured to better than 50%. Symbols are color-
coded according to the evolutionary state of the host star (see
Figure 7). TOI-197 b is highlighted with a orange/red/blue
star symbol. Magenta letters show the position of solar sys-
tem planets.
radius relations between sub-Saturns and Jupiters. We
note that TOI-197.01 is one of the most precisely char-
acterized Saturn-sized planets to date, with a density
uncertainty of ∼ 15%.
6. CONCLUSIONS
We have presented the discovery of TOI-197.01, the
first transiting planet orbiting an oscillating host star
identified by TESS. Our main conclusions are as follows:
101100101102103Orbital Period (Days)101100Planet Radius (Jupiter Radii)(a)DwarfsSubgiantsGiants101100101102103104Incident Flux (Earth Units)101100Planet Radius (Jupiter Radii)(b)DwarfsSubgiantsGiants103102101100101Planet Mass (Jupiter Masses)101100Planet Radius (Jupiter Radii)JSNEDwarfs SubgiantsGiants14
Huber et al.
• TOI-197 is a late subgiant / early red giant with
a clear presence of mixed modes. Combined
spectroscopy and asteroseismic modeling revealed
that the star has just started its ascent on the
red giant branch, with R(cid:63) = 2.943 ± 0.064R(cid:12),
M(cid:63) = 1.212 ± 0.074M(cid:12) and near-solar age
(4.9 ± 1.1 Gyr). TOI-197 is a typical oscillating
star expected to be detected with TESS, and
demonstrates the power of asteroseismology even
with only 27 days of data.
• The oscillation amplitude of TOI-197 is consistent
with ensemble measurements from Kepler . This
confirms that the redder bandpass of TESS com-
pared to Kepler only has a small effect on the os-
cillation amplitude (as expected from scaling re-
lations, Kjeldsen & Bedding 1995; Ballot et al.
2011a), supporting the expected yield of thou-
sands of solar-like oscillators with 2-minute ca-
dence observations in the nominal TESS mission
(Schofield et al. 2018). A detailed study of the
asteroseismic performance of TESS will have to
await ensemble measurements of noise levels and
amplitudes.
• TOI-197.01 is a "hot Saturn" (F = 343 ± 24F⊕,
Rp = 0.836 ± 0.031 RJ, Mp = 0.190 ± 0.018 MJ)
and joins a small but growing population of close-
in, transiting planets orbiting evolved stars. Based
on its incident flux, radius and mass, TOI-197.01
may be a precursor to the population of gas gi-
ants that undergo radius re-inflation due to the
increased irradiance as their host star evolves up
the red-giant branch.
• TOI-197.01 is one the most precisely character-
ized Saturn-sized planets to date, with a density
measured to ∼ 15%. TOI-197.01 does not follow
the trend of increasing planet mass with host star
metallicity discovered in sub-Saturns with sizes be-
tween 4−8 R⊕, which has been linked to metal-rich
disks preferentially forming more massive planet
cores (Petigura et al. 2017b). The moderate den-
sity (ρp = 0.431 ± 0.062 g cm−3) suggests that
Saturn-sized planets may follow a relatively nar-
row range of densities, a possible signature of the
transition in the interior structure leading to dif-
ferent mass-radius relations for sub-Saturns and
Jupiters.
TOI-197 provides a first glimpse at the strong poten-
tial of TESS to characterize exoplanets using asteroseis-
mology. TOI-197.01 has one the most precisely charac-
terized densities of known Saturn-sized planets to date,
with an uncertainty of ∼ 15%. Thanks to asteroseis-
mology the planet density uncertainty is dominated by
measurements of the transit depth and the radial ve-
locity amplitude, and thus can be expected to further
decrease with continued transit observations and radial
velocity follow-up, which is readily performed given the
brightness (V=8) of the star. Ensemble studies of such
precisely characterized planets orbiting oscillating sub-
giants can be expected to yield significant new insights
on the effects of stellar evolution on exoplanets, comple-
menting current intensive efforts to characterize planets
orbiting dwarfs.
The authors wish to recognize and acknowledge the
very significant cultural role and reverence that the
summit of Maunakea has always had within the indige-
nous Hawai'ian community. We are most fortunate to
have the opportunity to conduct observations from this
mountain. We thank Andrei Tokovinin for helpful in-
formation on the Speckle observations obtained with
SOAR. D.H. acknowledges support by the National
Aeronautics and Space Administration through the
TESS Guest Investigator Program (80NSSC18K1585)
and by the National Science Foundation (AST-1717000).
A.C. acknowledges support by the National Science
Foundation under the Graduate Research Fellowship
Program. W.J.C., W.H.B., A.M., O.J.H. and G.R.D.
acknowledge support from the Science and Technol-
ogy Facilities Council and UK Space Agency. H.K.
and F.G. acknowledge support from the European So-
cial Fund via the Lithuanian Science Council grant
No. 09.3.3-LMT-K-712-01-0103. Funding for the Stel-
lar Astrophysics Centre is provided by The Danish Na-
tional Research Foundation (Grant DNRF106). A.J. ac-
knowledges support from FONDECYT project 1171208,
CONICYT project BASAL AFB-170002, and by the
Ministry for the Economy, Development, and Tourism's
Programa Iniciativa Cient´ıfica Milenio through grant
IC 120009, awarded to the Millennium Institute of As-
trophysics (MAS). R.B. acknowledges support from
FONDECYT Post-doctoral Fellowship Project 3180246,
and from the Millennium Institute of Astrophysics
(MAS). A.M.S. is supported by grants ESP2017-82674-
R (MINECO) and SGR2017-1131 (AGAUR). R.A.G.
and L.B. acknowledge the support of the PLATO
grant from the CNES. The research leading to the
presented results has received funding from the Eu-
ropean Research Council under the European Com-
munity's Seventh Framework Programme (FP72007-
2013) ERC grant agreement no 338251 (StellarAges).
S.M. acknowledges support from the European Research
Council through the SPIRE grant 647383. This work
TOI-197
15
was also supported by FCT (Portugal) through na-
tional funds and by FEDER through COMPETE2020
by these grants: UID/FIS/04434/2013 & POCI-01-
0145-FEDER-007672, PTDC/FIS-AST/30389/2017 &
POCI-01-0145-FEDER-030389. T.L.C. acknowledges
support from the European Union's Horizon 2020 re-
search and innovation programme under the Marie
Sk(cid:32)lodowska-Curie grant agreement No. 792848 (PUL-
SATION). E.C.
is funded by the European Unions
Horizon 2020 research and innovation program un-
der the Marie Sklodowska-Curie grant agreement No.
664931. V.S.A. acknowledges support from the In-
dependent Research Fund Denmark (Research grant
7027-00096B). D.S. acknowledges support from the Aus-
tralian Research Council.
S.B. acknowledges NASA
grant NNX16AI09G and NSF grant AST-1514676.
T.R.W. acknowledges support from the Australian Re-
search Council through grant DP150100250. A.M. ac-
knowledges support from the ERC Consolidator Grant
funding scheme (project ASTEROCHRONOMETRY,
G.A. n.
772293). S.M. acknowledges support from
the Ramon y Cajal
fellowship number RYC-2015-
17697. M.S.L. is supported by the Carlsberg Foundation
(Grant agreement no.: CF17-0760). A.M. and P.R. ac-
knowledge support from the HBCSE-NIUS programme.
J.K.T. and J.T. acknowledge that support for this work
was provided by NASA through Hubble Fellowship
grants HST-HF2-51399.001 and HST-HF2-51424.001
awarded by the Space Telescope Science Institute, which
is operated by the Association of Universities for Re-
search in Astronomy, Inc., for NASA, under contract
NAS5-26555. T.S.R. acknowledges financial support
from Premiale 2015 MITiC (PI B. Garilli). This project
has been supported by the NKFIH K-115709 grant and
the Lendulet Program of the Hungarian Academy of
Sciences, project No. LP2018-7/2018.
Based on observations made with the Hertzsprung
SONG telescope operated on the Spanish Observatorio
del Teide on the island of Tenerife by the Aarhus and
Copenhagen Universities and by the Instituto de As-
trofsica de Canarias. Funding for the TESS mission is
provided by NASA's Science Mission directorate. We
acknowledge the use of public TESS Alert data from
pipelines at the TESS Science Office and at the TESS
Science Processing Operations Center. This research
has made use of the Exoplanet Follow-up Observation
Program website, which is operated by the California
Institute of Technology, under contract with the Na-
tional Aeronautics and Space Administration under the
Exoplanet Exploration Program. This paper includes
data collected by the TESS mission, which are publicly
available from the Mikulski Archive for Space Telescopes
(MAST).
Software: Astropy(AstropyCollaborationetal.2018),
Matplotlib (Hunter 2007), DIAMONDS (Corsaro & De
Ridder 2014), isoclassify (Huber et al. 2017), EXOFASTv2
(Eastman 2017), ktransit (Barclay 2018)
REFERENCES
Allard, F., Homeier, D., & Freytag, B. 2012, Royal Society
Barclay, T. 2018, ktransit: Exoplanet transit modeling tool
of London Philosophical Transactions Series A, 370,
2765, doi: 10.1098/rsta.2011.0269
in python, Astrophysics Source Code Library.
http://ascl.net/1807.028
Andr´e, Q., Barker, A. J., & Mathis, S. 2017, A&A, 605,
Barclay, T., Rowe, J. F., Lissauer, J. J., et al. 2013, Nature,
A117, doi: 10.1051/0004-6361/201730765
494, 452, doi: 10.1038/nature11914
Appourchaux, T., Chaplin, W. J., Garc´ıa, R. A., et al.
Bastien, F. A., Stassun, K. G., Basri, G., & Pepper, J.
2012, A&A, 543, A54, doi: 10.1051/0004-6361/201218948
2013, Nature, 500, 427, doi: 10.1038/nature12419
Astropy Collaboration, Price-Whelan, A. M., Sipocz, B. M.,
et al. 2018, AJ, 156, 123, doi: 10.3847/1538-3881/aabc4f
Basu, S., & Chaplin, W. J. 2017, Asteroseismic Data
Analysis: Foundations and Techniques
Baglin, A., Auvergne, M., Boisnard, L., et al. 2006, in 36th
Bazot, M., Vauclair, S., Bouchy, F., & Santos, N. C. 2005,
COSPAR Scientific Assembly, Vol. 36, 3749
A&A, 440, 615, doi: 10.1051/0004-6361:20052698
Ball, W. H., & Gizon, L. 2017, A&A, 600, A128,
Bedding, T. R. 2012, in Astronomical Society of the Pacific
doi: 10.1051/0004-6361/201630260
Ballard, S., Chaplin, W. J., Charbonneau, D., et al. 2014,
ApJ, 790, 12, doi: 10.1088/0004-637X/790/1/12
Conference Series, Vol. 462, Progress in Solar/Stellar
Physics with Helio- and Asteroseismology, ed.
H. Shibahashi, M. Takata, & A. E. Lynas-Gray, 195
Ballot, J., Barban, C., & van't Veer-Menneret, C. 2011a,
Bedding, T. R., Kjeldsen, H., Arentoft, T., et al. 2007, ApJ,
A&A, 531, 124, doi: 10.1051/0004-6361/201016230
663, 1315, doi: 10.1086/518593
Ballot, J., Gizon, L., Samadi, R., et al. 2011b, A&A, 530,
Bell, K. J., Hekker, S., & Kuszlewicz, J. S. 2019, MNRAS,
A97, doi: 10.1051/0004-6361/201116547
482, 616, doi: 10.1093/mnras/sty2731
16
Huber et al.
Benomar, O., Baudin, F., Chaplin, W. J., Elsworth, Y., &
Chontos, A., Huber, D., Latham, D. W., et al. 2019, AJ, in
Appourchaux, T. 2012, MNRAS, 420, 2178,
doi: 10.1111/j.1365-2966.2011.20184.x
press, arXiv:1903.01591.
https://arxiv.org/abs/1903.01591
Benomar, O., Masuda, K., Shibahashi, H., & Suto, Y. 2014,
Christensen-Dalsgaard, J. 2008, Ap&SS, 316, 13,
PASJ, doi: 10.1093/pasj/psu069
doi: 10.1007/s10509-007-9675-5
Berger, T. A., Huber, D., Gaidos, E., & van Saders, J. L.
Ciardi, D. R., Beichman, C. A., Horch, E. P., & Howell,
2018, ApJ, 866, 99, doi: 10.3847/1538-4357/aada83
Bodenheimer, P., Lin, D. N. C., & Mardling, R. A. 2001,
ApJ, 548, 466, doi: 10.1086/318667
Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science,
327, 977, doi: 10.1126/science.1185402
Bouchy, F., Bazot, M., Santos, N. C., Vauclair, S., &
Sosnowska, D. 2005, A&A, 440, 609,
doi: 10.1051/0004-6361:20052697
Bovy, J., Rix, H.-W., Green, G. M., Schlafly, E. F., &
Finkbeiner, D. P. 2016, ApJ, 818, 130,
doi: 10.3847/0004-637X/818/2/130
Brahm, R., Jord´an, A., & Espinoza, N. 2017a, PASP, 129,
034002, doi: 10.1088/1538-3873/aa5455
Brahm, R., Jord´an, A., Hartman, J., & Bakos, G. 2017b,
MNRAS, 467, 971, doi: 10.1093/mnras/stx144
Brahm, R., Espinoza, N., Jord´an, A., et al. 2018, arXiv
e-prints. https://arxiv.org/abs/1811.02156
Bressan, A., Marigo, P., Girardi, L., et al. 2012, MNRAS,
427, 127, doi: 10.1111/j.1365-2966.2012.21948.x
Buchhave, L. A., Latham, D. W., Johansen, A., et al. 2012,
Nature, 486, 375, doi: 10.1038/nature11121
Bugnet, L., Garc´ıa, R. A., Davies, G. R., et al. 2018, A&A,
620, A38, doi: 10.1051/0004-6361/201833106
Campante, T. L. 2018, Asteroseismology and Exoplanets:
Listening to the Stars and Searching for New Worlds, 49,
55, doi: 10.1007/978-3-319-59315-9 3
Campante, T. L., Lund, M. N., Kuszlewicz, J. S., et al.
2016a, ApJ, 819, 85, doi: 10.3847/0004-637X/819/1/85
Campante, T. L., Schofield, M., Kuszlewicz, J. S., et al.
2016b, ApJ, 830, 138, doi: 10.3847/0004-637X/830/2/138
Ceillier, T., van Saders, J., Garc´ıa, R. A., et al. 2016,
MNRAS, 456, 119, doi: 10.1093/mnras/stv2622
Chaplin, W. J., Elsworth, Y., Davies, G. R., et al. 2014a,
MNRAS, 445, 946, doi: 10.1093/mnras/stu1811
Chaplin, W. J., & Miglio, A. 2013, ARA&A, 51, 353,
doi: 10.1146/annurev-astro-082812-140938
S. B. 2015, ApJ, 805, 16,
doi: 10.1088/0004-637X/805/1/16
Claret, A., & Bloemen, S. 2011, A&A, 529, A75,
doi: 10.1051/0004-6361/201116451
Corsaro, E., & De Ridder, J. 2014, A&A, 571, A71,
doi: 10.1051/0004-6361/201424181
Corsaro, E., De Ridder, J., & Garc´ıa, R. A. 2015, A&A,
579, A83, doi: 10.1051/0004-6361/201525895
Corsaro, E., Frohlich, H.-E., Bonanno, A., et al. 2013,
MNRAS, 430, 2313, doi: 10.1093/mnras/stt059
Corsaro, E., Mathur, S., Garc´ıa, R. A., et al. 2017, A&A,
605, A3, doi: 10.1051/0004-6361/201731094
Creevey, O. L., Metcalfe, T. S., Schultheis, M., et al. 2017,
A&A, 601, A67, doi: 10.1051/0004-6361/201629496
Cunha, M. S., Stello, D., Avelino, P. P.,
Christensen-Dalsgaard, J., & Townsend, R. H. D. 2015,
ApJ, 805, 127, doi: 10.1088/0004-637X/805/2/127
Davies, G. R., & Miglio, A. 2016, Astronomische
Nachrichten, 337, 774, doi: 10.1002/asna.201612371
Davies, G. R., Silva Aguirre, V., Bedding, T. R., et al.
2016, MNRAS, 456, 2183, doi: 10.1093/mnras/stv2593
Deheuvels, S., & Michel, E. 2011, A&A, 535, A91,
doi: 10.1051/0004-6361/201117232
Demarque, P., Guenther, D. B., Li, L. H., Mazumdar, A., &
Straka, C. W. 2008, Ap&SS, 316, 31,
doi: 10.1007/s10509-007-9698-y
Demory, B.-O., & Seager, S. 2011, The Astrophysical
Journal Supplement Series, 197, 12,
doi: 10.1088/0067-0049/197/1/12
Eastman, J. 2017, EXOFASTv2: Generalized
publication-quality exoplanet modeling code,
Astrophysics Source Code Library.
http://ascl.net/1710.003
Eastman, J., Gaudi, B. S., & Agol, E. 2013, PASP, 125, 83,
doi: 10.1086/669497
Evans, D. W., Riello, M., De Angeli, F., et al. 2018, A&A,
616, A4, doi: 10.1051/0004-6361/201832756
Chaplin, W. J., Basu, S., Huber, D., et al. 2014b, ApJS,
Fur´esz, G. 2008, PhD thesis, University of Szeged, Szeged,
210, 1, doi: 10.1088/0067-0049/210/1/1
Chen, J., & Kipping, D. 2017, ApJ, 834, 17,
doi: 10.3847/1538-4357/834/1/17
Hungary
Furlan, E., Ciardi, D. R., Everett, M. E., et al. 2017, AJ,
153, 71, doi: 10.3847/1538-3881/153/2/71
Choi, J., Dotter, A., Conroy, C., et al. 2016, ApJ, 823, 102,
Garc´ıa, R. A., Hekker, S., Stello, D., et al. 2011, MNRAS,
doi: 10.3847/0004-637X/823/2/102
414, L6, doi: 10.1111/j.1745-3933.2011.01042.x
TOI-197
17
Garc´ıa, R. A., Ceillier, T., Salabert, D., et al. 2014, A&A,
572, A34, doi: 10.1051/0004-6361/201423888
Gilbert, J., Bergmann, C., Bloxham, G., et al. 2018, in
Society of Photo-Optical Instrumentation Engineers
(SPIE) Conference Series, Vol. 10702, Ground-based and
Airborne Instrumentation for Astronomy VII, 107020Y
Gilliland, R. L., McCullough, P. R., Nelan, E. P., et al.
2011, ApJ, 726, 2, doi: 10.1088/0004-637X/726/1/2
Jenkins, J. M., Twicken, J. D., McCauliff, S., et al. 2016, in
Proc. SPIE, Vol. 9913, Software and Cyberinfrastructure
for Astronomy IV, 99133E
Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp,
J. R. 2010, PASP, 122, 905, doi: 10.1086/655775
Johnson, J. A., Petigura, E. A., Fulton, B. J., et al. 2017,
AJ, 154, 108, doi: 10.3847/1538-3881/aa80e7
Jones, M. I., Brahm, R., Espinoza, N., et al. 2018, A&A,
Gizon, L., Ballot, J., Michel, E., et al. 2013, Proceedings of
613, A76, doi: 10.1051/0004-6361/201731478
the National Academy of Science, 110, 13267,
doi: 10.1073/pnas.1303291110
Grec, G., Fossat, E., & Pomerantz, M. A. 1983, SoPh, 82,
55
Grunblatt, S. K., Huber, D., Gaidos, E. J., et al. 2016, AJ,
152, 185, doi: 10.3847/0004-6256/152/6/185
Grunblatt, S. K., Huber, D., Gaidos, E., et al. 2017, AJ,
154, 254, doi: 10.3847/1538-3881/aa932d
-- . 2018, ApJL, 861, L5, doi: 10.3847/2041-8213/aacc67
Grundahl, F., Fredslund Andersen, M.,
Christensen-Dalsgaard, J., et al. 2017, ApJ, 836, 142,
doi: 10.3847/1538-4357/836/1/142
Guenel, M., Mathis, S., & Remus, F. 2014, A&A, 566, L9,
doi: 10.1051/0004-6361/201424010
Handberg, R., Brogaard, K., Miglio, A., et al. 2017,
MNRAS, 472, 979, doi: 10.1093/mnras/stx1929
Handberg, R., & Campante, T. L. 2011, A&A, 527, A56,
doi: 10.1051/0004-6361/201015451
Handberg, R., & Lund, M. N. 2014, MNRAS, 445, 2698,
doi: 10.1093/mnras/stu1823
Hekker, S., Elsworth, Y., De Ridder, J., et al. 2011, A&A,
525, A131, doi: 10.1051/0004-6361/201015185
Høg, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A,
355, L27
Hubbard, W. B., Burrows, A., & Lunine, J. I. 2002,
ARA&A, 40, 103,
doi: 10.1146/annurev.astro.40.060401.093917
Huber, D., Stello, D., Bedding, T. R., et al. 2009,
Communications in Asteroseismology, 160, 74.
https://arxiv.org/abs/0910.2764
Kallinger, T., Beck, P. G., Stello, D., & Garcia, R. A. 2018,
A&A, 616, A104, doi: 10.1051/0004-6361/201832831
Kallinger, T., Hekker, S., Garcia, R. A., Huber, D., &
Matthews, J. M. 2016, Science Advances, 2, 1500654,
doi: 10.1126/sciadv.1500654
Kallinger, T., Hekker, S., Mosser, B., et al. 2012, A&A,
541, A51, doi: 10.1051/0004-6361/201218854
Kaufer, A., Stahl, O., Tubbesing, S., et al. 1999, The
Messenger, 95, 8
Kjeldsen, H., & Bedding, T. R. 1995, A&A, 293, 87
Kjeldsen, H., Bedding, T. R., & Christensen-Dalsgaard, J.
2008a, ApJL, 683, L175, doi: 10.1086/591667
Kjeldsen, H., Bedding, T. R., Butler, R. P., et al. 2005,
ApJ, 635, 1281, doi: 10.1086/497530
Kjeldsen, H., Bedding, T. R., Arentoft, T., et al. 2008b,
ApJ, 682, 1370, doi: 10.1086/589142
Kov´acs, G., Bakos, G. ´A., Hartman, J. D., et al. 2010, ApJ,
724, 866, doi: 10.1088/0004-637X/724/2/866
Lainey, V., Jacobson, R. A., Tajeddine, R., et al. 2017,
Icarus, 281, 286, doi: 10.1016/j.icarus.2016.07.014
Lebreton, Y., & Goupil, M. J. 2014, A&A, 569, A21,
doi: 10.1051/0004-6361/201423797
Lenz, P., & Breger, M. 2005, Communications in
Asteroseismology, 146, 53, doi: 10.1553/cia146s53
Lopez, E. D., & Fortney, J. J. 2016, ApJ, 818, 4,
doi: 10.3847/0004-637X/818/1/4
Lund, M. N., Handberg, R., Davies, G. R., Chaplin, W. J.,
& Jones, C. D. 2015, ApJ, 806, 30,
doi: 10.1088/0004-637X/806/1/30
Huber, D., Bedding, T. R., Arentoft, T., et al. 2011, ApJ,
731, 94, doi: 10.1088/0004-637X/731/2/94
Huber, D., Chaplin, W. J., Christensen-Dalsgaard, J., et al.
2013a, ApJ, 767, 127, doi: 10.1088/0004-637X/767/2/127
Huber, D., Carter, J. A., Barbieri, M., et al. 2013b, Science,
342, 331. https://arxiv.org/abs/1310.4503
Lund, M. N., Handberg, R., Kjeldsen, H., Chaplin, W. J.,
& Christensen-Dalsgaard, J. 2017, in European Physical
Journal Web of Conferences, Vol. 160, European Physical
Journal Web of Conferences, 01005
Lund, M. N., Lundkvist, M., Silva Aguirre, V., et al. 2014,
A&A, 570, A54, doi: 10.1051/0004-6361/201424326
Huber, D., Bryson, S. T., Haas, M. R., et al. 2017, ApJ,
Lundkvist, M. S. 2015, PhD thesis, Stellar Astrophysics
224, 2, doi: 10.3847/0067-0049/224/1/2
Centre, Aarhus University, Denmark
Hunter, J. D. 2007, Computing In Science & Engineering,
Lundkvist, M. S., Kjeldsen, H., Albrecht, S., et al. 2016,
9, 90, doi: 10.1109/MCSE.2007.55
Iben, Jr., I. 1965, ApJ, 142, 1447, doi: 10.1086/148429
Nature Communications, 7, 11201,
doi: 10.1038/ncomms11201
18
Huber et al.
Mann, A. W., Newton, E. R., Rizzuto, A. C., et al. 2016,
Quinn, S. N., White, T. R., Latham, D. W., et al. 2015,
AJ, 152, 61, doi: 10.3847/0004-6256/152/3/61
ApJ, 803, 49, doi: 10.1088/0004-637X/803/2/49
Mardling, R. A. 2011, in European Physical Journal Web of
Rayner, J., Bond, T., Bonnet, M., et al. 2012, in
Conferences, Vol. 11, European Physical Journal Web of
Conferences, 03002
Proc. SPIE, Vol. 8446, Ground-based and Airborne
Instrumentation for Astronomy IV, 84462C
Mathur, S., Garc´ıa, R. A., R´egulo, C., et al. 2010, A&A,
Ricker, G. R., Winn, J. N., Vanderspek, R., et al. 2014, in
511, A46, doi: 10.1051/0004-6361/200913266
Mathur, S., Hekker, S., Trampedach, R., et al. 2011, ApJ,
741, 119, doi: 10.1088/0004-637X/741/2/119
Mayor, M., Pepe, F., Queloz, D., et al. 2003, The
Messenger, 114, 20
Miller, N., & Fortney, J. J. 2011, ApJL, 736, L29,
doi: 10.1088/2041-8205/736/2/L29
Morel, P., & Lebreton, Y. 2008, Ap&SS, 316, 61,
doi: 10.1007/s10509-007-9663-9
Mosser, B., Elsworth, Y., Hekker, S., et al. 2012a, A&A,
537, A30, doi: 10.1051/0004-6361/201117352
Mosser, B., Goupil, M. J., Belkacem, K., et al. 2012b,
A&A, 548, A10, doi: 10.1051/0004-6361/201220106
Mosser, B., Benomar, O., Belkacem, K., et al. 2014, A&A,
572, L5, doi: 10.1051/0004-6361/201425039
Mosumgaard, J. R., Ball, W. H., Silva Aguirre, V., Weiss,
A., & Christensen-Dalsgaard, J. 2018, MNRAS, 478,
5650, doi: 10.1093/mnras/sty1442
Nielsen, L. D., Bouchy, F., Turner, O., et al. 2019, A&A,
623, A100, doi: 10.1051/0004-6361/201834577
Ogilvie, G. I. 2014, ARA&A, 52, 171,
doi: 10.1146/annurev-astro-081913-035941
Ong, J. M. J., & Basu, S. 2019, ApJ, 870, 41,
doi: 10.3847/1538-4357/aaf1b5
Pande, D., Bedding, T. R., Huber, D., & Kjeldsen, H. 2018,
MNRAS, 480, 467, doi: 10.1093/mnras/sty1869
Paxton, B., Bildsten, L., Dotter, A., et al. 2011, ApJS, 192,
3, doi: 10.1088/0067-0049/192/1/3
Paxton, B., Cantiello, M., Arras, P., et al. 2013, ApJS, 208,
4, doi: 10.1088/0067-0049/208/1/4
Paxton, B., Marchant, P., Schwab, J., et al. 2015, ApJS,
220, 15, doi: 10.1088/0067-0049/220/1/15
Pesnell, W. D. 1990, ApJ, 363, 227, doi: 10.1086/169333
Petigura, E. 2015, PhD Thesis, University of California.
https://arxiv.org/abs/1510.03902
Petigura, E. A., Howard, A. W., Marcy, G. W., et al.
2017a, AJ, 154, 107, doi: 10.3847/1538-3881/aa80de
Petigura, E. A., Sinukoff, E., Lopez, E. D., et al. 2017b, AJ,
153, 142, doi: 10.3847/1538-3881/aa5ea5
Society of Photo-Optical Instrumentation Engineers
(SPIE) Conference Series, Vol. 9143, , 20
Rodrigues, T. S., Girardi, L., Miglio, A., et al. 2014,
MNRAS, 445, 2758, doi: 10.1093/mnras/stu1907
Rodrigues, T. S., Bossini, D., Miglio, A., et al. 2017,
MNRAS, 467, 1433, doi: 10.1093/mnras/stx120
Roxburgh, I. W. 2017, A&A, 604, A42,
doi: 10.1051/0004-6361/201731057
Schlaufman, K. C., & Winn, J. N. 2013, ApJ, 772, 143,
doi: 10.1088/0004-637X/772/2/143
Schlegel, D. J., Finkbeiner, D. P., & Davis, M. 1998, ApJ,
500, 525, doi: 10.1086/305772
Schofield, M., Chaplin, W. J., Huber, D., et al. 2018, ApJS,
submitted
Scuflaire, R., Montalb´an, J., Th´eado, S., et al. 2008,
Ap&SS, 316, 149, doi: 10.1007/s10509-007-9577-6
Serenelli, A., Johnson, J., Huber, D., et al. 2017, ApJS,
233, 23, doi: 10.3847/1538-4365/aa97df
Silva Aguirre, V., Davies, G. R., Basu, S., et al. 2015,
MNRAS, 452, 2127, doi: 10.1093/mnras/stv1388
Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ,
131, 1163, doi: 10.1086/498708
Sliski, D. H., & Kipping, D. M. 2014, ApJ, 788, 148,
doi: 10.1088/0004-637X/788/2/148
Smith, J. C., Stumpe, M. C., Van Cleve, J. E., et al. 2012,
PASP, 124, 1000, doi: 10.1086/667697
Sousa, S. G. 2014, ARES + MOOG: A Practical Overview
of an Equivalent Width (EW) Method to Derive Stellar
Parameters, ed. E. Niemczura, B. Smalley, & W. Pych,
297 -- 310
Sousa, S. G., Adibekyan, V., Delgado-Mena, E., et al. 2018,
A&A, 620, A58, doi: 10.1051/0004-6361/201833350
Stassun, K. G., & Torres, G. 2016, AJ, 152, 180,
doi: 10.3847/0004-6256/152/6/180
-- . 2018, ApJ, 862, 61, doi: 10.3847/1538-4357/aacafc
Stassun, K. G., Oelkers, R. J., Pepper, J., et al. 2018, AJ,
156, 102, doi: 10.3847/1538-3881/aad050
Steffen, J. H., Ragozzine, D., Fabrycky, D. C., et al. 2012,
Proceedings of the National Academy of Science, 109,
7982, doi: 10.1073/pnas.1120970109
Pires, S., Mathur, S., Garc´ıa, R. A., et al. 2015, A&A, 574,
Stello, D., Zinn, J., Elsworth, Y., et al. 2017, ApJ, 835, 83,
A18, doi: 10.1051/0004-6361/201322361
doi: 10.3847/1538-4357/835/1/83
Queloz, D., Mayor, M., Udry, S., et al. 2001, The
Stephan, A. P., Naoz, S., & Gaudi, B. S. 2018, AJ, 156,
Messenger, 105, 1
128, doi: 10.3847/1538-3881/aad6e5
TOI-197
19
Stumpe, M. C., Smith, J. C., Catanzarite, J. H., et al. 2014,
Verner, G. A., Elsworth, Y., Chaplin, W. J., et al. 2011,
PASP, 126, 100, doi: 10.1086/674989
Tayar, J., & Pinsonneault, M. H. 2018, ApJ, 868, 150,
doi: 10.3847/1538-4357/aae979
Tayar, J., Stassun, K. G., & Corsaro, E. 2018, arXiv
e-prints. https://arxiv.org/abs/1812.04010
Thorngren, D. P., & Fortney, J. J. 2018, AJ, 155, 214,
doi: 10.3847/1538-3881/aaba13
Tokovinin, A., Mason, B. D., & Hartkopf, W. I. 2010, AJ,
139, 743, doi: 10.1088/0004-6256/139/2/743
MNRAS, 415, 3539,
doi: 10.1111/j.1365-2966.2011.18968.x
Villaver, E., Livio, M., Mustill, A. J., & Siess, L. 2014, ApJ,
794, 3, doi: 10.1088/0004-637X/794/1/3
Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, in
Society of Photo-Optical Instrumentation Engineers
(SPIE) Conference Series, Vol. 2198, Society of
Photo-Optical Instrumentation Engineers (SPIE)
Conference Series, ed. D. L. Crawford & E. R. Craine,
362
Vrard, M., Mosser, B., Barban, C., et al. 2015, A&A, 579,
Torres, G., Fischer, D. A., Sozzetti, A., et al. 2012, ApJ,
A84, doi: 10.1051/0004-6361/201425064
757, 161, doi: 10.1088/0004-637X/757/2/161
Wang, S., Jones, M., Shporer, A., et al. 2019, AJ, 157, 51,
Townsend, R. H. D., & Teitler, S. A. 2013, MNRAS, 435,
3406, doi: 10.1093/mnras/stt1533
Van Eylen, V., & Albrecht, S. 2015, ApJ, 808, 126,
doi: 10.1088/0004-637X/808/2/126
Van Eylen, V., Dai, F., Mathur, S., et al. 2018, MNRAS,
478, 4866, doi: 10.1093/mnras/sty1390
Van Eylen, V., Albrecht, S., Huang, X., et al. 2019, AJ,
157, 61, doi: 10.3847/1538-3881/aaf22f
doi: 10.3847/1538-3881/aaf1b7
Weiss, L. M., Marcy, G. W., Rowe, J. F., et al. 2013, ApJ,
768, 14, doi: 10.1088/0004-637X/768/1/14
Weiss, W. W., Moffat, A. F. J., & Kudelka, O. 2008,
Communications in Asteroseismology, 157, 271
White, T. R., Bedding, T. R., Stello, D., et al. 2011, ApJ,
743, 161, doi: 10.1088/0004-637X/743/2/161
Wittenmyer, R. A., Endl, M., Wang, L., et al. 2011, ApJ,
743, 184, doi: 10.1088/0004-637X/743/2/184
Wizinowich, P., Acton, D. S., Shelton, C., et al. 2000,
van Saders, J. L., Ceillier, T., Metcalfe, T. S., et al. 2016,
PASP, 112, 315, doi: 10.1086/316543
Nature, 529, 181, doi: 10.1038/nature16168
van Saders, J. L., & Pinsonneault, M. H. 2013, ApJ, 776,
67, doi: 10.1088/0004-637X/776/2/67
Vanzi, L., Zapata, A., Flores, M., et al. 2018, MNRAS, 477,
5041, doi: 10.1093/mnras/sty936
Veras, D. 2016, Royal Society Open Science, 3, 150571,
doi: 10.1098/rsos.150571
Wright, E. L., Eisenhardt, P. R. M., Mainzer, A. K., et al.
2010, AJ, 140, 1868, doi: 10.1088/0004-6256/140/6/1868
Yıldız, M., C¸ elik Orhan, Z., & Kayhan, C. 2016, MNRAS,
462, 1577, doi: 10.1093/mnras/stw1709
Yu, J., Huber, D., Bedding, T. R., & Stello, D. 2018,
MNRAS, 480, L48, doi: 10.1093/mnrasl/sly123
Zapolsky, H. S., & Salpeter, E. E. 1969, ApJ, 158, 809,
doi: 10.1086/150240
|
1708.00693 | 1 | 1708 | 2017-08-02T10:42:06 | WASP-12b: A Mass-Losing Extremely Hot Jupiter | [
"astro-ph.EP"
] | WASP-12b is an extreme hot Jupiter in a 1 day orbit, suffering profound irradiation from its F type host star. The planet is surrounded by a translucent exosphere which overfills the Roche lobe and produces line-blanketing absorption in the near-UV. The planet is losing mass. Another unusual property of the WASP-12 system is that observed chromospheric emission from the star is anomalously low: WASP-12 is an extreme outlier amongst thousands of stars when the log $R^{'}_{HK}$ chromospheric activity indicator is considered. Occam's razor suggests these two extremely rare properties coincide in this system because they are causally related. The absence of the expected chromospheric emission is attributable to absorption by a diffuse circumstellar gas shroud which surrounds the entire planetary system and fills our line of sight to the chromospherically active regions of the star. This circumstellar gas shroud is probably fed by mass loss from WASP-12b. The orbital eccentricity of WASP-12b is small but may be non-zero. The planet is part of a hierarchical quadruple system; its current orbit is consistent with prior secular dynamical evolution leading to a highly eccentric orbit followed by tidal circularization. When compared with the Galaxy's population of planets, WASP-12b lies on the upper boundary of the sub-Jovian desert in both the $(M_{\rm P}, P)$ and $(R_{\rm P}, P)$ planes. Determining the mass loss rate for WASP-12b will illuminate the mechanism(s) responsible for the sub-Jovian desert. | astro-ph.EP | astro-ph | WASP-12b: A Mass-Losing Extremely Hot
Jupiter
Carole A. Haswell
School of Physical Sciences, The Open University, Walton Hall, Milton Keynes, MK7
6AA, United Kingdom
e-mail: [email protected]
Text finalised June 2017. Submitted 2nd August 2017.
Abstract
WASP-12b is an extreme hot Jupiter in a 1 day orbit, suffering profound irradiation from
its F type host star. The planet is surrounded by a translucent exosphere which overfills the
Roche lobe and produces line-blanketing absorption in the near-UV. The planet is losing
mass. Another unusual property of the WASP-12 system is that observed chromospheric
emission from the star is anomalously low: WASP-12 is an extreme outlier amongst
thousands of stars when the log R'HK chromospheric activity indicator is considered.
Occam's razor suggests these two extremely rare properties coincide in this system because
they are causally related. The absence of the expected chromospheric emission is
attributable to absorption by a diffuse circumstellar gas shroud which surrounds the entire
planetary system and fills our line of sight to the chromospherically active regions of the
star. This circumstellar gas shroud is probably fed by mass loss from WASP-12b. The
orbital eccentricity of WASP-12b is small but may be non-zero. The planet is part of a
hierarchical quadruple system; its current orbit is consistent with prior secular dynamical
evolution leading to a highly eccentric orbit followed by tidal circularization. When
compared with the Galaxy's population of planets, WASP-12b lies on the upper boundary
of the sub-Jovian desert in both the (MP, P) and (RP, P) planes. Determining the mass loss
rate for WASP-12b will illuminate the mechanism(s) responsible for the sub-Jovian desert.
Introduction
In 1995 our preconceptions about planetary systems were challenged with the discovery of
51 Peg b, the first hot Jupiter (Mayor and Queloz 1995). The subject of this Chapter,
WASP-12b, is one of the hottest of the hot Jupiters, and its extreme properties have
encouraged many studies. This contribution focuses on the phenomenon of mass loss from
this extreme planet and on its implications for the exoplanet population, without attempting
to be an exhaustive review of other properties. WASP-12b was an early discovery by the
SuperWASP planet search; its large radius, RP = 1.7 RJ, and short orbital period, P=1.09 d,
both contribute to favourable probabilities for detection by transit (see e.g. Haswell 2010
for details.) WASP-12b orbits an F star (Hebb et al 2009), which, along with the very short
orbital period, means it is among the most irradiated of the known exoplanets. The planet
mass is MP = 1.4 MJ and the orbital semi-major axis is 0.023 AU. Of these planet
parameters, as so often in astronomy, the period is by far the best determined. The other
planet parameters rest upon our knowledge of the basic stellar parameters, and their
1
reported uncertainties often do not completely propagate our uncertainties about the host
star. Haswell (2010) gives an explanation of the derivation of the fundamental parameters
for transiting planets.
The host star of this extreme hot Jupiter, WASP-12, has effective temperature Teff =
6250±100 K and surface gravity log g = 4.2± 0.2 and is metal rich compared to the Sun; the
star's age is between 1 and 2.65 Gyr and its mass M* is between 1.23Mʘ and 1.49 Mʘ
(Fossati et al 2010a, Hebb et al 2009). The planet's orbit places it only about 1.5 stellar
radii above the star's photosphere, making it an extreme hot Jupiter. Swain et al (2013)
found that WASP-12b's near-IR brightness temperature is around 3000K, approximately
twice that of the archetypal hot Jupiters HD 189733b and HD 209458b. The distance to
WASP-12 is less precisely known, with Fossati et al (2010a) finding a possible range of
295-465 pc. These limits are based on the star's securely measured effective temperature
and surface gravity; narrower uncertainty ranges have been quoted elsewhere, but these
ranges may rest upon debatable assumptions. See Fossati et al (2010a) for a detailed
discussion of WASP-12's fundamental parameters, including the distance.
The seminal UV observations of the first known transiting exoplanet, HD 209458b, by
Vidal-Madjar et al (2003), showed that hot Jupiter planets are surrounded by extensive
exospheres. HD 209458b produces a 15% deep transit in the Lyman α line, indicating the
planet is surrounded by an exosphere with radius > 3 RP and containing neutral hydrogen.
UV observations underpin the study of exoplanet exospheres because this spectral region
contains strong resonance lines of many common species. Strong resonance lines have
large oscillator strengths and involve the ground state and can thus be detected even when
the emitting or absorbing column density is modest. Consequently, these lines are (i)
prominent in the emission from stellar chromospheres and (ii) provide a particularly
sensitive probe for the presence of diffuse gas surrounding a planet. Taken together, this
means the UV contains many features which are particularly useful for transmission
spectroscopy of transiting planets.
UV Observations of WASP-12b transits
It was immediately clear upon the discovery of WASP-12b that this system is a good
candidate to search for exospheric gas surrounding the planet. The short orbital period
combined with the inflated planet radius mean that the planet fills a substantial fraction of
its Roche lobe. WASP-12b remains one of the three known giant planets with the largest
Roche lobe filling factor (Busuttil 2017). Hubble Space Telescope (HST) observations were
promptly proposed after the discovery. The obvious goal was to perform transmission
spectroscopy covering Lyman α to determine the extent of the HI exosphere for such a
strongly irradiated planet. This proved impractical because WASP-12 is 6 to 10 times more
distant than HD 209458b and consequently the predicted stellar photon count-rate for
WASP-12 was too low. HD 209458b at 47 pc is close enough for a precise distance
determination by Hipparcos parallax measurements, so produced good far UV count-rates
despite there being so little stellar flux in the far-UV region which contains Lyman α. To
overcome the distance of WASP-12 and the intrinsic faintness of FGK stars in the far UV,
2
the first HST observations of WASP-12 pioneered HST transmission spectroscopy in the
near-UV wavelength range, λλ 2539 – 2829 Å (Haswell et al 2012) where the underlying
stellar photospheric flux is enhanced by a factor of ~106. This proved to be an extremely
informative wavelength region. Despite this, it has remained under-exploited in subsequent
studies of exoplanet exospheres.
Hubble Space Telescope Cosmic Origins Spectrograph observations of WASP-12b
The observations presented in Haswell et al (2012) are a superset of those presented in
Fossati et al (2010b) and used the slitless Cosmic Origins Spectrograph (COS) to obtain 10
HST orbits of R~20,000 near-UV spectroscopy. COS records data in three wavelength
regions known as 'stripes' A, B and C. This data was obtained on two distinct HST visits,
each centred on a transit of WASP-12b; the visits were timed so that the orbital coverage of
WASP-12b interleaves. The two visits used slightly different grating settings, so that each
of the 3 stripes only partially overlaps in wavelength between the two visits. The
significantly noiser data subsequently obtained and reported by Nichols et al (2015) used
exclusively the Visit 1 setting of Haswell et al (2012).
The near-UV spectral region contains thousands of overlapping photospheric absorption
lines. The strongest of these are resonance lines, including the Mg II h&k lines, Mn II
2577Å and Fe II 2586Å which can be easily picked out in Figure 17 of Haswell et al
(2012). Unevolved stars generally exhibit chromospheric emission in the line cores of these
features. The plethora of weaker lines blend to produce substantial stellar photospheric
absorption of the stellar continuum which illuminates the bottom of the stellar photosphere.
This line-blanketing stellar photospheric absorption is strongest in the C stripe, and weakest
but still approaching 50% in the B stripe, as shown in Figure 1 of Haswell et al (2012).
Photometric Analysis
Because slightly different wavelength regions were covered, it is impossible to produce a
set of homogeneous light curves including all the photons collected in the observations
reported in Haswell et al (2012). The comprehensive analysis is described in detail in the
original paper, the broad conclusions are summarised here.
In both visits, the transit of WASP-12b in the near-UV region is deeper than that of the
opaque planet, indicating that the planet is surrounded by an exosphere of strongly
absorbing gas. In the first visit, reported by Fossati et al (2010b), the relative depths of the
transits in the A, B, and C stripes corresponds roughly to the relative amounts of absorption
in these regions within the stellar atmosphere. The near-UV-absorbing gas thus appears to
have a temperature and composition similar to those of a stellar photosphere. This finding
was independently corroborated by the interpretation of IR transmission spectroscopy and
secondary eclipse data by Swain et al (2013). The transit depth in the A and C stripes of
Visit 1 of Haswell et al (2012) is 3.5%, which makes it clear that the absorbing gas
surrounding WASP-12b overfills the planet's Roche lobe, at least some of the time. In turn,
3
this implies that WASP-12b is losing mass.
In detail, the repeated observations reported in Haswell et al (2012) and Nichols et al
(2015) do not reproduce the shape of the Visit 1 transit light curve. In particular the early
ingress for Visit 1 reported in Fossati et al (2010b) is not a general feature. Visit 2 of
Haswell et al (2012) instead shows a high point at the phase of optical ingress: Visit 2 and
the four visits of Nichols et al (2015) suggest the absorbing gas is patchy, and its spatial
coverage as seen from our vantage point extends to the earliest observed orbital phases, at
around φ=0.83. Figure 9 of Carroll-Nellenback et al (2017) visualises the dispersed
material lost in a hydrodynamic outflow from a hot Jupiter. It shows a patchy distribution
of material which extends azimuthally to surround the star. The patchy and variable
absorption observed in the near-UV lightcurves of WASP-12, as described in Haswell et al
(2012) and Nichols et al (2015), is consistent with this picture.
Spectroscopic Analysis
Despite the extension of absorbing gas spread azimuthally around the orbit of WASP-12b,
there is a consistent relative overdensity around the planet itself. There is clear evidence for
this: the near-UV light curves consistently show fluxes during transit which are depressed
by more than the 1.5% dip caused by the opaque planet itself (Fossati et al 2010b, Haswell
et al 2012, Nichols et al 2015). This means that the unocculted light from the star suffers
from increased absorption at phases where the planet is in transit compared to that at other
observed phases. This can be seen for individual lines, too: there is evidence from the very
strong Mg II h&k lines that there is some absorption from diffuse gas at all observed phases
(see below), but despite this we can perform transmission spectroscopy by comparing
spectra obtained during transit with those obtained away from transit. The comparison
allows us to identify excess absorption attributable to the higher column density of
intervening diffuse gas during the planet transit.
Fossati et al (2010b) performed transmission spectroscopy of WASP-12b for Visit 1 of
Haswell et al (2012) finding enhanced transit depths at the wavelengths of resonance lines
of neutral sodium, tin, and manganese, and of singly ionised ytterbium, scandium,
manganese, aluminium, vanadium and magnesium. Haswell et al (2012) detected repeated
enhanced absorption during transit in 65 distinct features. These features include exospheric
absorption throughout the inner wings of the very strong Mg II h&k lines on both visits and
a detection of exospheric absorption in Fe II 2586Å, which remains the heaviest species
detected in an exoplanet transit. Absorption at these individual features is accompanied by
line-blanketing absorption which accumulates to produce the broad-band absorption seen in
the near-UV light curves. Not all the features in the WASP-12b transmission spectrum can
be unambiguously identified (Haswell et al 2012). In particular, in some cases it is unclear
whether a feature is due to a resonance line of a rare element or a weaker feature of a more
abundant element, or a blend of the two. These cases should be revisited in the future when
we have a better understanding of the composition of WASP-12b's outer layers and the
mechanisms underlying the mass loss which feeds the exospheric gas.
4
WASP-12's missing chromospheric emission in MgII h&k
All the observations of WASP-12's Mg II h&k lines show zero flux in the line cores
(Fossati et al 2010b, Haswell et al 2012, Nichols et al 2015). The strong Fe II 2586Å line,
covered only in Visit 2 of Haswell et al 2012, is also consistent with zero flux in the line
core. This makes WASP-12's observed spectrum unique: all other main sequence or
slightly evolved stars of WASP-12's spectral type show chromospheric emission in these
features. A basal level of chromospheric flux is seen even for old, slowly rotating, inactive
stars of this type.
Occam's razor suggests that this anomaly in WASP-12's spectrum is related to the extreme
exoplanet it hosts. Haswell et al (2012) examined the possibility that the Mg II h&k line
core flux is absorbed by the interstellar medium (ISM), concluding that an anomalously
over-dense line of sight would be required. The missing chromospheric emission in Mg II
h&k and the Fe II 2586Å line is more likely to be absorbed locally by a diffuse
circumstellar gas shroud fed by the mass-loss from the overflowing exosphere of the
extreme hot Jupiter planet WASP-12b.
Stellar Chromospheric emission shrouded by planetary mass loss
The Mg II h&k lines can only be observed from space because their wavelengths are in the
UV. However, the analogous resonance lines of the singly ionised element from the next
period in the periodic table, Ca, are in the optical and are accessible from the ground.
Consequently, the Ca II H&K lines (λλ 3968Å, 3933Å) are commonly-used indicators of
stellar chromospheric activity, usually parameterised in terms of log R'HK (Noyes et al
1984). This particular metric is useful because it allows intercomparison of the activity
levels of F, G, and K type stars on a conveniently calibrated scale; see Staab et al (2017),
their Figures 2 & 7, for an illustration of how log R'HK is defined. In particular, for a main
sequence FGK star, log R'HK is expected to exceed a value of -5.1 which corresponds to a
basal level of activity observed in the oldest, least active, slowest rotating stars of this type.
The basal emission persists even when stars are completely devoid of active regions
(Schröder et al 2012). WASP-12 lies significantly below this limit with log R'HK= -5.5, a
value determined by Knutson et al (2010) who did not remark upon its anomalous nature.
There has been significant interest in the measured log R'HK values of transiting planet host
stars as there appears to be a correlation between the host star's log R'HK value and the
planet's surface gravity (Hartman 2010, Figueira et al 2014, Fossati, Ingrassia & Lanza
2015). Fossati et al (2013) and Staab et al (2017) compared log R'HK for WASP-12 and
other planet hosts with an extensive sample of field stars. While Staab et al (2017) show
that a quarter of known host stars of close-in transiting planets show anomalously low Ca II
H&K chromospheric emission, WASP-12 remains the most extreme outlier.
Fossati et al (2013) and Staab et al (2017) interpret the missing Ca II H&K chromospheric
emission in these close-in planet host stars as due to absorption by diffuse gas shrouding
these planetary systems. This builds on and corroborates the conclusions Haswell et al
5
(2012) drew from WASP-12's missing Mg II h&k and Fe II chromospheric emission. In
each of these low log R'HK systems, gas originates in an outflow from the close-in ablating
planet, forming a diffuse cloud of circumstellar gas which shrouds the entire planetary
system. It is plausible that the extremely irradiated planet WASP-12b has a more
prodigious mass outflow than more moderately irradiated giant planets. While the detailed
dynamics of these outflows is largely unexplored, it is natural to expect this to lead to a
higher column density of gas in the line of sight to the chromospherically active regions of
this particular star.
Significantly, Staab et al (2017) show that two low-mass close-in planet hosts, Kepler-25
and Kepler-28, have anomalously low values of log R'HK, establishing that the phenomenon
does not require the presence of a gas giant planet. Indeed the most observationally
spectacular examples of ablating planets are the catastrophically disintegrating exoplanets
(CDEs), exemplified by Kepler 1520b. The CDEs are extremely close-in planets, detected
via their periodic but variable transits. The transits are attributed to a cloud of dust formed
by material ablated from the rocky surface of a low mass planet irradiated to temperature
T~2100K. In the case of Kepler 1520b, the planet is thought to have a mass of about
0.1M⊕. At temperatures of ~2100K rocky minerals sublime, and a thermal wind of metal-
rich vapour and entrained dust flows from the planet. The entrained dust can vary in spatial
distribution and optical depth, thus producing variable transits (Rappaport et al 2012). The
transiting dust cloud model was confirmed when Bochinski et al (2015) detected the colour
dependence of transit depth of Kepler 1520b. The known CDE host stars are sadly too
distant and faint to encourage transmission spectroscopy with current facilities.
WASP-12b and Mass Loss in Hot Jupiter Planets
Since the discovery of the extended HI exosphere of HD 209458b there has been significant
interest in the phenomenon of mass loss in hot Jupiter planets. In this section we will
discuss first the models, then the relationship between the models and WASP-12b, which is
the most extreme known hot Jupiter, and finally we will consider the potential future
observations of WASP-12b which might clarify some of the presently open questions.
Mass-loss rates in hot Jupiters: brief overview of the state of the art
The discovery of the extended HI exosphere of HD 209458b stimulated prompt modelling
work to calculate atmospheric loss from hot Jupiters (e.g. Lammer et al 2003). There has
been much activity in this area in the intervening years, with recent papers including
Jackson et al (2017) and Carroll-Nellenback et al (2017). Lammer et al (2003) showed that
irradiation by stellar X-ray and extreme-UV (EUV) flux leads to atmospheric expansion,
and mass-loss rates of ≈1012 g s-1 for HD 209458b. Later work (e.g. Erkaev et al 2007)
incorporated the Roche potential, which lowers the energy required for atmospheric escape.
In the extreme case of a Roche lobe-filling planet, there is no energy barrier to atmospheric
escape through the saddle point at L1. Recent models including Tripathi et al (2015) and
Jackson et al (2017) variously incorporate Roche lobe overflow, the evolutionary response
6
of the planet to mass loss, the irradiative heating and the three-dimensional hydrodynamic
outflow. Computational limitations mean that no single model has yet simultaneously
included all of these ingredients self-consistently, but the crucial factors which determine
the mass-loss rate for close-in planets are being established and quantitatively tested.
As laid out in Frank et al (2002), the mass loss rate depends critically on the comparision
between the evolution of the mass donor's size and the evolution of the Roche lobe size.
This in turn depends on the transfer of orbital angular momentum within the system.
Valsecchi et al (2015) investigated tidally driven Roche lobe overflow of hot Jupiters
within this framework, including the effects of irradiation and planet evolution. They
produce evolutionary tracks tracing the mass, radius, and orbital period for 1 MJ planets
orbiting 1Mʘ stars, both for conservative and non-conservative mass transfer. In
conservative mass transfer all mass leaving the planet is accreted onto the star; conversely,
non-conservative mass transfer means some fraction of the mass from the planet leaves the
system. In non-conservative mass transfer, angular momentum will generally be lost.
Observational evidence in WASP-12 and other irradiated planet host stars for the presence
of circumstellar gas shrouds (Haswell et al 2012, Fossati et al 2013, Staab et al 2017)
implies that mass transfer is non-conservative. This has implications for the planets' orbital
evolution, which transit timing measurements over long temporal baselines can test.
The case of WASP-12b
As noted before, WASP-12b is among the most irradiated of exoplanets; temperatures in
the outer layers of the planet's atmosphere are in the same regime as M dwarf star
photospheres. It is an extremely hot Jupiter. As shown in Figure 21 of Haswell et al (2012)
WASP-12b is close to filling its Roche lobe, with a volume filling fraction of 0.54 (Busuttil
2017). A planet in this configuration is expected to lose mass through hydrodynamic
escape. Stellar irradiation will photoionize the upper atmosphere, with the liberated
electrons causing collisional heating. This heating drives a hydrodynamic wind comprised
of the constituents of the upper atmospheric gas (c.f. Erkaev et al 2007, Sanz-Forcada et al
2010, Tripathi et al 2015; Jackson et al 2017.) Bisikalo et al (2013) performed a
hydrodynamic simulation of outflows from both the L1 and L2 points of WASP-12b's
Roche lobe, concluding that a steady-state gaseous envelope encompassing the Roche lobe
can result. This simulation did not include treatment of the photoionisation which drives the
wind, instead setting a priori a temperature of 104K outside the Roche lobe. More recent
work by Carroll-Nellenback et al (2017) has taken a more self-consistent approach to the
dynamics of similar 'up-orbit' and 'down-orbit' outflows, though not tailored to the
specific case of WASP-12b.
Husnoo et al (2011), Albrecht et al (2012) and Collins et al (2017) analysed radial velocity
data of WASP-12 from SOPHIE in which the Rossiter McLaughlin effect is of comparable
amplitude to the systematic instrumental effects. Thus the spin of WASP-12 appears to be
slow or approximately orthogonal to the orbital angular momentum of WASP-12b's orbit.
For a slowly-spinning host star under the assumptions of Valsecchi et al (2015), tides
7
transfer angular momentum from the orbit to the stellar spin, causing orbital decay. For
WASP-12b, there is evidence that the orbit may be tidally shrinking in this way
(Maciejewski et al 2016), but see also Collins et al (2017).
All these considerations imply that WASP-12b is likely to be shedding mass more rapidly
than (almost?) all other known exoplanets. This is consistent with the detection of gas
surrounding the planet during transit, which presents an effective obscuring area in the
near-UV of up to three times that of the optically opaque planet (Haswell et al 2012), and in
excess of the cross-section presented by the planet's Roche lobe. The anomalously
depressed stellar chromospheric emission from WASP-12 at all observed orbital phases is
also consistent with dispersed gas shrouding the entire planetary system. Haswell et al
(2012), Fossati et al (2013) and Staab et al (2017) show that WASP-12 has the lowest
chromospheric emission of any star measured, which suggests that either the intrinsic
stellar activity is extremely low, WASP-12b's mass loss is prodigious and produces a
higher column depth of shrouding circumstellar gas than is present in any other system, or
both. WASP-12b may be in the early stages of the evolutionary phase discussed by
Valsecchi et al (2014). They suggest Roche lobe-filling hot Jupiters may lose their entire
gaseous envelopes, leaving hot super-Earths.
The 'up-orbit' stream of Carroll-Nellenback et al (2017) can be accelerated away from the
star due to ram pressure from the stellar wind. The simulations neglect radiation pressure:
the absorption of stellar emission found by Fossati et al (2010b), Haswell et al (2012) and
Nichols et al (2015) guarantees radiation pressure will tend to drive WASP-12's
circumstellar gas outwards.
The quantitative self-consistent numerical calculations of mass loss rates for particular hot
Jupiter systems have generally been tailored to more moderately irradiated planets at larger
orbital distances. Bisikalo et al (2013) did not calculate a mass loss rate for WASP-12b.
The value of mass loss rate of 10-7 MJ year-1 calculated by Li et al (2010) for WASP-12b
was predicated on an erroneous non-zero eccentricity inferred from the discovery paper
radial velocity data (Hebb et al 2009). Husnoo et al (2011) subsequently derived e=0.017
+0.015 -0.010, i.e. an approximately circular orbit. Lai et al (2010) obtained a mass loss rate 25
times lower with assumptions of an isothermal planet atmosphere of 3000K, a density at L1
of 3×1012 g cm-3 and a nozzle for Roche lobe overflow which is of linear size of 0.6 planet
radii. It is fair to say not all of these assumptions are supported in detail. For example,
using atmospheric structure models motivated by the observations of Swain et al (2013)
and Mandell et al (2013) can change the mass loss rate by four orders of magnitude. Thus
the mass loss rate for WASP-12b remains undetermined.
Observations to determine the mass loss rate of WASP-12b
The flux of ionising stellar radiation upon the planet is a crucial parameter for calculations
of the photoevaporative mass loss from hot Jupiters. For stars of WASP-12's spectral type
and late type stars generally, this flux largely arises from stellar activity. The Mg II and
Ca II lines indicate that WASP-12's apparent chromospheric activity is anomalously low,
but unless the star is unique, this must be due to absorption of the intrinsic chromospheric
8
emission. The ionising flux could in principle be directly measured, but WASP-12 is too
distant for this to be practical, because ionising flux is strongly absorbed by the ISM (see
Fossati et al 2013 for a general discussion of the effects of interstellar absorption on the
observed chromospheric emission from exoplanet host stars). An alternative approach
would be to take a spectrum in the far-UV (i.e around 1100- 2500 Å) as Fossati et al (2015)
have done for WASP-13. This wavelength region contains strong chromospheric emission
lines of ionised species including C IV and Si IV. Because these species are rare in the
ISM, these lines are not absorbed even for moderate hydrogen column densities. Linsky et
al (2013, 2014) observed nearby stars and constructed empirical relationships between the
various components of stellar chromospheric emission. Using this, one can take the
C IV and Si IV line fluxes and infer the intrinsic fluxes of the remainder of the
chromospheric emission. Thus for WASP-13, Fossati et al (2015) were able to derive the
intrinsic stellar EUV flux, and subsequently model the upper atmosphere of WASP-13b,
deriving a mass-loss rate of 1.5 ×1011 g s-1.
A good signal to noise spectrum of WASP-12 in the 1100- 2500 Å wavelength region will
require a substantial investment of time on HST or a successor UV space telescope. Given
the prodigious recent and ongoing modelling activity, this could be a proportionate and
appropriate investment with widespread implications.
Optical light curves measuring the transits of WASP-12b are much easier to obtain. Even
relatively small telescopes can determine transit timings to useful precision. Maciejewski et
al (2016) and Collins et al (2017) analyse optical transit timings, using 31 (23) transit
lightcurves spanning Nov 2012 – Feb 2016 (Nov 2009 – Feb 2015) respectively. Both
analyses incorporate additional orbital constraints from radial velocity (RV) measurements
and secondary eclipse timings, with the analysis of Maciejewski et al (2016) allowing a
finite eccentricity in their fit while Collins et al (2017) fix the eccentricity to zero. Collins
et al (2017) conclude that the transit timings are consistent with a linear ephemeris, while
Maciejewski et al (2016) present evidence for deviations. Maciejewski et al (2016) fit two
models: a quadratic ephemeris consistent with a simple tidal decay, and a 3300d periodic
signal attributed to the periastron precession of a slightly eccentric orbit with e=0.00110±
0.00036. Data from winter 2016/17 should distinguish between these two models. Bonomo
et al (2017) used transit timings, secondary eclipse timings, all suitable RVs from the
literature published up to 1 January 2016 and 15 unpublished HARPS-N RV points taken
before February 2015 to perform an Bayesian determination of WASP-12b's orbital
parameters, finding e < 0.020 (1σ upper limit). Thus, the published data is consistent with a
circular orbit for WASP-12b. Knowledge of the current orbital configuration is a vital
ingredient to determine the tidal effects on the planet which in turn strongly affect the
internal heating and mass loss rate, as dramatically illustrated by Li et al (2010)'s work
which assumed a much larger eccentricity. The present orbital configuration can also
potentially rule out entire classes of evolutionary scenarios. This in turn has potentially
profound implications for the planet structure resulting from the mass loss history, and
consequently on the present evolution of the planet radius in response to mass loss. See
Valsecchi et al (2015) for modelling work which encapsulates such considerations.
Further high quality radial velocity measurements covering the phases of the WASP-12b
transit should allow the stellar spin to be better constrained. The existing SOPHIE
9
observations and analysis (Husnoo et al 2011, Albrecht et al 2012 and Collins et al 2017)
show scattered residuals with deviations from the fit comparable in amplitude to the
Rossiter McLaughlin effect itself. The superior precision and stability of HARPS-N should
allow for a much more precise determination of the projected stellar spin-orbit angle. This
information can be used to perform calculations of the tidal transfer of angular meomentum
from the orbit to the stellar spin, using techniques similar to those of Valsecchi et al (2015).
Testing the Pollution Hypothesis
In the currently accepted interpretation of the evidence we have discussed for the WASP-12
system, some fraction of the mass lost from WASP-12b is dispersed outwards, and is
probably carrying angular momentum. In the terminology of Frank et al (2002) and
Valsecchi et al (2015) this is an example of non-conservative mass transfer. A simple
analysis of mass transfer in the Roche geometry reveals that the minimum energy
configuration which satisfies the conservation laws is to move a small amount of mass to
very large distances. This mass carries the angular momentum away and allows the bulk of
the remaining mass to settle into the deepest part of the Roche potential. If this captures the
fate of the gas lost from WASP-12b, then the bulk of the material may have accreted onto
the star. Fossati et al (2010b) performed a detailed analysis to examine the element by
element photospheric abundances of WASP-12 for signs of accretion of material from the
planet. Some hints of this atmospheric pollution were found, but a more detailed
differential analysis comparing WASP-12 to stellar twins is required for a definitive
answer. This would be analogous to the differential abundance analyses of Melendez et al
(2009) on solar twins and Nickson (2015) on 61 Virginis and closely matched comparison
stars.
Implications for exoplanet demographics
The historical collection of biological specimens led to an appreciation of the relationships
between various species, and ultimately to our understanding of evolution and the
expression of genetic characteristics encoded in DNA. Similarly, the purpose behind this
exoplanets fauna is to illuminate the relationships between the Galaxy's planetary systems,
using well-studied exemplars to reveal the long-timescale evolutionary processes at work.
WASP-12b is a useful exemplar of a giant planet suffering high insolation as a result of a
very short orbital period around a hot (F-type) host star. It provides an interesting and
informative test-case for the mechanisms leading to one of the most prominent features of
the emerging demographic information about the Galaxy's population of planets.
The Sub-Jovian Desert
When only 106 transiting exoplanets were known, Szabó & Kiss (2011) identified a very
prominent feature in the distribution of planet radii with orbital period. There is a distinct
lack of planets with radii between about 2 R⊕ and 1RJ for orbital periods shorter than about
10
2.5 days. This feature has subsequently been called the Neptunian desert or the sub-Jovian
desert by various authors. As more transiting planets were discovered the feature persisted,
as seen, e.g., in Figure 4 of Mazeh et al (2016) showing the RP vs P distribution for almost
4500 transiting planets and candidate planets from Kepler. With the addition of thousands
more objects, there are a few within the sub-Jovian desert in the (RP, P) plane. In contrast
the (MP, P) plane shown in Figure 1 of Mazeh et al (2016) includes only the best-studied
1037 transiting planets. In this Figure the sub-Jovian desert is remarkably well-defined.
The sub-Jovian desert cannot be attributed to selection biases: short period planets are
favoured by both the radial velocity and the transit detection methods. The former favours
massive planets, while the latter favours large planets, and with both methods planets of
low mass and / or small radius have been detected below the desert. See, e.g., Haswell
(2010) for a discussion of these selection biases.
Mazeh et al (2016) perform a statistical analysis of subsets of the populations they consider,
allowing them to derive algebraic forms for the sharply defined upper boundary and the less
distinct lower boundary. They find
and
𝑀𝑃
𝑀𝐽
𝑅𝑃
𝑅𝐽
≅ (1.7 ± 0.2) (
𝑃
−1.14±0.14
)
𝑑𝑎𝑦
≅ (1.4 ± 0.3) (
𝑃
0.31±0.12
)
𝑑𝑎𝑦
for the upper boundary. WASP-12b itself was considered in the subset contributing to the
derivation of the boundary in the (MP, P) plane, but not in the (RP, P) plane where it lay
below the period range used. We can use the well-determined orbital period of WASP-12b
to calculate the values of mass and radius predicted for a planet of this period which lies on
the boundary of the sub-Jovian desert. Comparing these values with the less-precisely-
determined empirical mass and radius for WASP-12b (see the introduction) we find that
within the uncertainties, WASP-12b lies on both these boundaries, though it lies about 1σ
on the low-mass, large radius side of the boundaries. Given that we see WASP-12b is
currently losing mass, this is intriguing though inconclusive.
Models to explain the sub-Jovian Desert
There have been a number of mechanisms proposed to explain the observed sub-Jovian
desert. The root cause(s) are not yet agreed, and it is possible that multiple mechanisms
contribute to the explanation. Since WASP-12b lies at or near the well-defined upper
boundary, we will focus predominantly on the explanations for this feature.
Mass loss due to the prodigious insolation suffered by short orbital period planets is an
obvious candidate mechanism. Szabó & Kiss (2011) include this as the first of four possible
hypotheses to explain their findings, in particular for planets with loosely gravitationally
bound atmospheres. Indeed the correlation between surface gravity and orbital period found
by Southworth et al (2007) and Southworth (2008) is in fact the first discovery of the upper
11
boundary of the sub-Jovian desert.
Within the framework of planetary migration, Mazeh et al (2016) mention the possibility
that the upper boundary acts as a death line: planets which migrate to shorter orbital period
and cross this line lose the majority of their mass as a result of irradiation or Roche lobe-
overflow. Upon crossing the death line, planets would move rapidly downwards in the (MP,
P) and (RP, P) planes.
An alternative possibility sketched by Mazeh et al (2016) within the migration framework
instead invokes a correlation between protoplanetary disk mass and planet mass. The disk
pushes the planet inwards until the disk density diminishes below that required. More
massive protoplanetary discs might persist to longer times and / or smaller distances from
the star, thus growing more massive planets and pushing them further inwards.
Matsakos and Königl (2016) explain both the upper and lower boundaries simultaneously
by considering the fate of planets which arrive at short orbital periods via eccentric orbits
and subsequent circularisation. Planets can acquire an eccentric orbit with a small
pericentre either via planet-planet scattering or more gradual processes such as Kozai-
Lidov migration. If the pericentre of the eccentric orbit is less than the Roche limit 𝑎𝑅, the
simplest form of which is
𝑎𝑅 = 𝑅𝑃 (
1
3
)
2𝑀∗
𝑀𝑃
the planet will be tidally disrupted and will not survive to be subsequently discovered in a
circular orbit. Consequently there is a minimum orbital radius after circularisation which
depends on the planet's Roche limit. Since the Roche limit depends on the planet's mass
and radius, this leads to different orbital criteria for surviving giant and terrestrial planets.
Matsakos and Königl (2016) elegantly show that these criteria neatly lead to the observed
upper and lower boundaries for the sub-Jovian desert.
Most recently Ginzburg and Sari (2017) treat the Roche lobe overflow of hot Jupiters
analytically, and conclude that the sub-Jovian desert is a consequence of core masses > 6
M⊕ for the Galaxy's population of hot Jupiters.
WASP-12b confronts the models
To explain WASP-12b within Matsakos and Königl (2016)'s scenario, we need a precursor
phase in which the planet acquired a high eccentricity low pericentre orbit. Since WASP-12
is the primary star of a hierarchical triple star system (Bechter et al 2014) this can naturally
be explained by the secular dynamical evolution. Hamers and Lai (2017) present a
simplified semi-analytic model of this evolution, concluding that extremely high
eccentricities can be generated in the (WASP-12b) planetary orbit in this scenario. Note
that the alternative planet-planet scattering route to a high eccentricity orbit for WASP-12b
could be problematic because planet-planet scattering events are expected for only the first
<108 years, at least ten times less than the age of WASP-12.
12
Kurokawa and Nakamoto (2014) consider the effects of XUV-driven atmospheric escape
and Roche-lobe overflow on the exoplanet population. Their model assumes an insolation
appropriate to a G type host star and WASP-12b rests comfortably above their minimum
survival mass line for a planet with a 10M⊕ core. Clearly WASP-12b is losing mass, which
can be reconciled with this model if the (unknown) XUV radiation flux on the planet
exceeds that assumed and / or WASP-12b's mass loss rate is low enough that the planet
will persist at essentially the same mass. Currently, as we have seen, the mass loss rate for
WASP-12b is non-zero, but beyond that is very poorly constrained.
Conclusions
WASP-12b is a fascinating planet, and has several extreme properties. It has a short orbital
period, suffers very high insolation from its hot (F type) host star, and is among the largest
known planets. The planet's large size is probably related to its proximity to the star,
through irradiation and its highly filled Roche lobe. The optically opaque planet sits within
the Roche lobe, but is surrounded by an exosphere of translucent gas which overfills the
Roche lobe and produces line-blanketing absorption in the near-UV. The observed
chromospheric emission from the host star WASP-12 is anomalously low: it is an extreme
outlier in the distribution of the stellar chromospheric activity indicator log R'HK. We
attribute the anomalously depressed stellar chromospheric emission to absorption by a
diffuse shroud of circumstellar gas which fills our line of sight, covering the
chromospherically active regions of the star at all orbital phases. We surmise that the
circumstellar gas shroud originates in mass loss from the planet, via the exosphere.
The mass loss rate from WASP-12b is unknown. Establishing a high precision
measurement of the orbital eccentricity and monitoring the orbital period evolution over the
coming years and decades are needed to predict and confront estimates of the mass loss
rate. Observations which can achieve this include radial velocity measurements, transit
timings, and secondary eclipse timings. An empirical determination of the ionising EUV
flux experienced by WASP-12b is a further vital parameter. The best way to establish this
is with a UV spectrum covering the C IV and Si IV features. Meanwhile the various
ingredients needed to build computational models of the physical mechanisms underlying
mass loss from an extreme planet like WASP-12b exist, but have not yet been fully
integrated in a self-consistent way and applied to this case.
WASP-12b lies on the upper boundary of the sub-Jovian desert. Its position may be a
natural consequence, through secular dynamical evolution, of the planet's status as the
lowest mass object in a hierarchical quadruple system. It is unclear whether its current
evolution will move it rapidly down in mass and radius through the desert; nor is the
current and future orbital period evolution clear. Determination of the mass loss rate from
WASP-12b will clarify the evolutionary pathway(s) leading to one of the most prominent
demographic features of the Galaxy's population of exoplanets.
Acknowledgements
CH gratefully acknowledges financial support from STFC under grants ST/L000776/1 and ST/P000584/1,
and thanks her many primary literature co-authors for sharing their knowledge, skills and enthusiasm. Special
thanks to Joel and Charlotte for their forbearance while this Chapter was written.
13
References
Albrecht S, Winn JN, Johnson JA et al (2012) Obliquities of Hot Jupiter Host Stars: Evidence for Tidal
Interactions and Primordial Misalignments. ApJ 757:18
Bechter EB, Crepp JR, Ngo H et al (2014) WASP-12b and HAT-P-8b are Members of Triple Star Systems.
ApJ 788:2
Bisikalo D, Kaygorodov P, Ionov D et al (2013) Three-dimensional Gas Dynamic Simulation of the
Interaction between the Exoplanet WASP-12b and its Host Star. ApJ 764:19
Bochinski JJ, Haswell CA, Marsh TA et al (2015) Direct Evidence for an Evolving Dust Cloud from the
Exoplanet KIC 12557548 b. ApJLett 800:L21
Bonomo AS, Desidera S, Benatti S et al (2017) arXiv:1704.00373
Busuttil R (2017) Investigating Exoplanets and Transients Using Small-Aperture Telescopes. PhD Thesis,
The Open University
Carroll-Nellenback J, Frank A, Liu B et al (2017) Hot planetary winds near a star: dynamics, wind-wind
interactions, and observational signatures. MNRAS 466:2458-2473 DOI 10.1093/mnras/stw3307
Collins KA, Kielkopf JF, Stassun, KG (2017) Transit Timing Variation Measurements of WASP-12b and
Qatar-1b: No Evidence Of Additional Planets. AJ 153:78–91, DOI 10.3847/1538-3881/153/2/78,
1512.00464
Erkaev NV, Kulikov Yu N, Lammer, H et al (2007) Roche lobe effects on the atmospheric loss from 'Hot
Jupiters'. A&A 472:329
Figueira P, Oshagh M, Adibekyan VZh et al (2014) Revisiting the correlation between stellar activity and
planetary surface gravity, A&A 572, 51
Fossati L, Ayres TR, Haswell CA et al (2013) Absorbing Gas around the WASP-12 Planetary System.
ApJLett 766:20
Fossati L, Bagnulo S, Elmasli A et al (2010a) A Detailed Spectropolarimetric Analysis of the Planet-hosting
Star WASP-12. ApJ 720:872-886, DOI 10.1088/0004-637X/720/1/872
Fossati L, France K, Koskinen T et al (2015) Far-UV Spectroscopy of the Planet-hosting Star WASP-13:
High-energy Irradiance, Distance, Age, Planetary Mass-loss Rate, and Circumstellar Environment. ApJ
815:118
Fossati L, Haswell CA, Froning CS et al (2010b) Metals in the Exosphere of the Highly Irradiated Planet
WASP-12b. ApJLett 714:222-227, DOI 10.1088/2041-8205/714/2/L222
Fossati L, Ingrassia S & Lanza AF (2015) A Bimodal Correlation between Host Star Chromospheric
Emission and the Surface Gravity of Hot Jupiters. ApJLett 812, 35
Frank J, King A, Raine DJ (2002) Accretion Power in Astrophysics: Third Edition ISBN 0521620538.
Cambridge, UK: Cambridge University Press
Ginzburg, S Sari R (2017) Hot-Jupiter core mass from Roche lobe overflow MNRAS468:278-285
Hamers AS, Lai D (2017) Secular chaotic dynamics in hierarchical quadruple systems, with applications to
hot Jupiters in stellar binaries and triples. MNRAS in press. DOI: 10.1093/mnras/stx1319
Hartman JJ (2010) A Correlation Between Stellar Activity and the Surface Gravity of Hot Jupiters. ApJ 717
L138
Haswell CA (2010) Transiting Exoplanets. CUP Cambridge ISBN: 9780521139380
Haswell CA, Fossati L, Ayres T et al (2012) Near-ultraviolet Absorption, Chromospheric Activity, and Star-
Planet Interactions in the WASP-12 system. ApJ 760:A79, DOI 10.1088/0004-637X/760/1/79
Hebb L et al (2009) WASP-12b: The Hottest Transiting Extrasolar Planet Yet Discovered ApJ 693:1920,
DOI: 10.1088/0004-637X/693/2/1920
Husnoo N, Pont F, Hébrard G et al (2011) Orbital eccentricity of WASP-12 and WASP-14 from new radial
velocity monitoring with SOPHIE. MNRAS 413:2500
Jackson B, Arras P, Penev K et al (2017) A New Model of Roche Lobe Overflow for Short-period Gaseous
Planets and Binary Stars. ApJ 835:145
Knutson HA, Howard AW, Isaacson, H (2010) A Correlation Between Stellar Activity and Hot Jupiter
Emission Spectra ApJ 720:1569
Kurokawa H, Nakamoto T (2014) Mass-loss Evolution of Close-in Exoplanets: Evaporation of Hot Jupiters
and the Effect on Population ApJ 783:54
Lai D, Helling Ch, van den Heuvel EPJ (2010) Mass Transfer, Transiting Stream, and Magnetopause in
Close-in Exoplanetary Systems with Applications to WASP-12. ApJLett 721:L923
Lammer H, Selsis F, Ribas, I et al (2003) Atmospheric Loss of Exoplanets Resulting from Stellar X-Ray and
14
Extreme-Ultraviolet Heating. ApJLett 598:L121
Li S-L, Miller N, Lin DNC, Fortney JJ (2010) WASP-12b as a prolate, inflated and disrupting planet from
tidal dissipation Nature 463:1054
Linsky JL, France K, Ayres T (2013) Computing Intrinsic LYα Fluxes of F5 V to M5 V Stars. ApJ 766:69
Linsky JL, France K, Ayres T (2014) The Intrinsic Extreme Ultraviolet Fluxes of F5 V TO M5 V Stars. ApJ
780:61
Maciejewski G, Dimitrov D, Fernández M et al (2016) Departure from the constant-period ephemeris for the
transiting exoplanet WASP-12. A&A 588:L6
Mandell AM, Haynes K, Sinukoff E et al (2013) Exoplanet Transit Spectroscopy Using WFC3: WASP-12 b,
WASP-17 b, and WASP-19 b. ApJ 779:128
Matsakos T, Königl A (2016) On the Origin of the Sub-Jovian Desert in the Orbital-period-Planetary-mass
Plane ApJLett 820:L8
Mazeh T, Holczer T, Faigler S (2016) Dearth of short-period Neptunian exoplanets: A desert in period-mass
and period-radius planes A&A 589:75
Mayor M, Queloz D (1995) A Jupiter-mass companion to a solar-type star. Nature 378:355-359, DOI
10.1038/378355a0
Meléndez J, Asplund M, Gustafsson B, Yong D (2009) The Peculiar Solar Composition and Its Possible
Relation to Planet Formation. ApJLett 704:L66
Nichols JD; Wynn GA; Goad M, et al (2015) Hubble Space Telescope Observations of the NUV Transit of
WASP-12b. ApJ 890:133
Nickson, E (2015) Galactic chemical evolution signatures in stellar atmospheres and their possible planetary
connection. PhD Thesis, The Open University
Noyes RW, Hartmann LW, Baliunas SL, et al (1984) Rotation, convection, and magnetic activity in lower
main-sequence stars. ApJ 279:763
Rappaport S, Levine A, Chiang E et al (2012) Possible Disintegrating Short-period Super-Mercury Orbiting
KIC 12557548. ApJ 752:1 DOI: 10.1088/0004-637X/752/1/1
Sanz-Forcada J, Ribas I, Micela G et al (2010) A scenario of planet erosion by coronal radiation. A&A 511
L8
Schröder K-P, Mittag M, Pérez Martínez MI, et al (2012) Basal chromospheric flux and Maunder Minimum-
type stars: the quiet-Sun chromosphere as a universal phenomenon. A&A 540:130
Southworth J, Wheatley PJ, Sams G (2007) A method for the direct determination of the surface gravities of
transiting extrasolar planets. MNRAS 379:L11
Southworth J (2008) Homogeneous studies of transiting extrasolar planets - I. Light-curve analyses. MNRAS
386:1644
Staab D, Haswell CA, Smith GD et al (2017) SALT observations of the chromospheric activity of transiting
planet hosts: mass-loss and star-planet interactions. MNRAS 466:738 DOI: 10.1093/mnras/stw3172
Swain M, Deroo P, Tinetti G et al (2013) Probing the extreme planetary atmosphere of WASP-12b. Icarus
255:432, DOI: 10.1016/j.icarus.2013.04.003
Szabó Gy M, Kiss LL (2011) A Short-period Censor of Sub-Jupiter Mass Exoplanets with Low Density ApJ
727:44
Tripathi A, Kratter KM, Murray-Clay RA et al (2015) Simulated Photoevaporative Mass Loss from Hot
Jupiters in 3D. ApJ 808:173
Valsecchi F, Rappaport S, Rasio FA et al (2015) Tidally-driven Roche-lobe Overflow of Hot Jupiters with
MESA. ApJ 813:101
Valsecchi F, Rasio FA, Steffen JH (2014) From Hot Jupiters to Super-Earths via Roche Lobe Overflow.
ApJLett 793:L3
Vidal-Madjar A, Lecavelier des Etangs A, Désert J.-M. et al (2003) An extended upper atmosphere around
the extrasolar planet HD209458b. Nature 422:143-146, DOI 10.1038/nature01448
15
|
1802.00460 | 2 | 1802 | 2018-04-30T21:04:26 | OSSOS: X. How to use a Survey Simulator: Statistical Testing of Dynamical Models Against the Real Kuiper Belt | [
"astro-ph.EP"
] | All surveys include observational biases, which makes it impossible to directly compare properties of discovered trans-Neptunian Objects (TNOs) with dynamical models. However, by carefully keeping track of survey pointings on the sky, detection limits, tracking fractions, and rate cuts, the biases from a survey can be modelled in Survey Simulator software. A Survey Simulator takes an intrinsic orbital model (from, for example, the output of a dynamical Kuiper belt emplacement simulation) and applies the survey biases, so that the biased simulated objects can be directly compared with real discoveries. This methodology has been used with great success in the Outer Solar System Origins Survey (OSSOS) and its predecessor surveys. In this chapter, we give four examples of ways to use the OSSOS Survey Simulator to gain knowledge about the true structure of the Kuiper Belt. We demonstrate how to statistically compare different dynamical model outputs with real TNO discoveries, how to quantify detection biases within a TNO population, how to measure intrinsic population sizes, and how to use upper limits from non-detections. We hope this will provide a framework for dynamical modellers to statistically test the validity of their models. | astro-ph.EP | astro-ph | Draft version September 17, 2018
Typeset using LATEX preprint style in AASTeX62
8
1
0
2
r
p
A
0
3
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
0
6
4
0
0
.
2
0
8
1
:
v
i
X
r
a
OSSOS: X. How to use a Survey Simulator: Statistical Testing of Dynamical Models Against the
Real Kuiper Belt
S. M. Lawler,1 JJ. Kavelaars,1, 2 M. Alexandersen,3 M. T. Bannister,4 B. Gladman,5
J.-M. Petit,6 and C. Shankman2, 1, 7
1NRC-Herzberg Astronomy and Astrophysics, National Research Council of Canada, Victoria, BC, Canada
2Department of Physics and Astronomy, University of Victoria, Victoria, BC, Canada
3Institute of Astronomy and Astrophysics, Academia Sinica, Taipei, Taiwan
4Astrophysics Research Centre, Queen's University Belfast, Belfast, UK
5Department of Physics and Astronomy, The University of British Columbia, Vancouver, BC, Canada
6Institut UTINAM, UMR 6213 CNRS-Universit´e de Franche Comt´e, Besan¸con, France
7current affiliation: City of Toronto, ON, Canada
ABSTRACT
All surveys include observational biases, which makes it impossible to directly com-
pare properties of discovered trans-Neptunian Objects (TNOs) with dynamical models.
However, by carefully keeping track of survey pointings on the sky, detection limits,
tracking fractions, and rate cuts, the biases from a survey can be modelled in Survey
Simulator software. A Survey Simulator takes an intrinsic orbital model (from, for
example, the output of a dynamical Kuiper belt emplacement simulation) and applies
the survey biases, so that the biased simulated objects can be directly compared with
real discoveries. This methodology has been used with great success in the Outer Solar
System Origins Survey (OSSOS) and its predecessor surveys. In this chapter, we give
four examples of ways to use the OSSOS Survey Simulator to gain knowledge about
the true structure of the Kuiper Belt. We demonstrate how to statistically compare
different dynamical model outputs with real TNO discoveries, how to quantify detection
biases within a TNO population, how to measure intrinsic population sizes, and how
to use upper limits from non-detections. We hope this will provide a framework for
dynamical modellers to statistically test the validity of their models.
1. INTRODUCTION
The orbital structure, size frequency distribution and total mass of the trans-Neptunian region of
the Solar System is an enigmatic puzzle. Fernandez (1980) described an expected distribution for this
region based on the mechanisms for the delivery of cometary material into the inner Solar System.
Even before the first Kuiper belt object after Pluto was discovered, (1992 QB1; Jewitt & Luu 1993,
1995), it was theorized that dynamical effects produced by the mass contained in this region could
in principle be detectable (Hamid et al. 1968). The first discoveries made it clear that extracting
Corresponding author: S. M. Lawler
[email protected]
2
Lawler et al.
precise measurements of the orbital and mass distributions from this zone of the Solar System would
require careful analysis.
Major puzzles in the Solar System's history can be explored if one has accurate knowledge of
the distribution of material in this zone. Examples include: the orbital evolution of Neptune (e.g.
Malhotra 1993), the large scale re-ordering of the Solar System (e.g. Thommes et al. 1999; Gomes
et al. 2005), the process of planetesimal accretion (e.g. Stern 1996; Davis & Farinella 1997), the
production of cometary size objects via collisional processes (e.g. Stern 1995) and their delivery into
the inner Solar System (Duncan et al. 1988), and the stellar environment in which the Sun formed
(e.g. Brunini & Fernandez 1996; Kobayashi & Ida 2001). Our goal as observers is to test these
models and their consequences by comparison to the Solar System as we see it today. Given the
sparse nature of the datasets and the challenges of detecting and tracking trans-Neptunian objects
(TNOs), a strong statistical framework is required if we are to distinguish between these various
models.
The presence of large-scale biases in the detected sample of TNOs has been apparent since the
initial discoveries in the Kuiper belt, and multiple approaches have been used to account for these
biases. Jewitt & Luu (1995) use Monte-Carlo comparisons of Kuiper belt models to their detected
sample to estimate the total size of the Kuiper belt, taking into account the flux limits of their
survey. Similarly Irwin et al. (1995) estimate the flux limits of their searches and use these to weight
their detections and, combining those with the results reported in Jewitt & Luu (1995), provide an
estimate of the luminosity function of the region. Gladman et al. (1998) provide a Bayesian-based
analysis of their detected sample, combined with previously published surveys, to further refine the
measurement of the luminosity function of the Kuiper belt. Trujillo et al. (2001) determined the
size, inclination and radial distributions of the Kuiper belt by weighting the distribution of observed
TNOs based on their detectability and the fraction of the orbits that were contained within the
survey fields. Bernstein et al. (2004) refined the maximum-likelihood approach when they extended
the measurement of the size distribution to smaller scales and determined statistically significant
evidence of a break in the shape of the Kuiper belt luminosity function, later developed further by
the deeper survey of Fraser & Kavelaars (2009). A similar approach is taken in Adams (2010) who
make estimates of the underlying sampling by inverting the observed distributions. Other recent
results for Kuiper belt subpopulations include Schwamb et al. (2009), who use Monte Carlo sampling
to estimate detectability of Sedna-like orbits, and Parker (2015), who uses an approximate Bayesian
computation approach to account for unknown observation biases in the Neptune Trojans. Each of
these methods relies on backing out the underlying distributions from a detected sample.
Carefully measuring the true, unbiased structure of the Kuiper belt provides constraints on exactly
how Neptune migrated through the Kuiper belt. Two main models of Neptune's migration have been
proposed and modelled extensively. Pluto's eccentric, resonant orbit was first explained by a smooth
migration model for Neptune (Malhotra 1993, 1995). Larger scale simulations (Hahn & Malhotra
2005) showed this to be a viable way to capture many TNOs into Neptune's mean-motion resonances.
The so-called "Nice model" was proposed as an alternate way to destroy the proto-Kuiper belt and
capture many TNOs into resonances (Levison et al. 2008). In this model, the giant planets undergo
a dynamical instability that causes Neptune to be chaotically scattered onto an eccentric orbit that
damps to its current near-circular orbit while scattering TNOs and capturing some into its wide
resonances (Tsiganis et al. 2005). Due to the chaotic nature of this model, reproducing simulations
OSSOS: X. How to use a Survey Simulator
3
is difficult and many variations on the Nice model exist (e.g. Batygin et al. 2012; Nesvorn´y et al.
2013). One very recent and promising variation on the Nice model scenario includes the gravitational
effects of fairly large (∼Pluto-sized) bodies that cause Neptune's migration to be "grainy," having
small discrete jumps as these larger bodies are scattered (Nesvorn´y & Vokrouhlick´y 2016). More
dramatically, even larger planetary-scale objects could have transited and thus perturbed the young
Kuiper belt (Petit et al. 1999; Gladman & Chan 2006; Lykawka & Mukai 2008; Silsbee & Tremaine
2017)
The level of detail that must be included in Neptune migration scenarios is increasing with the
number of discovered TNOs with well-measured orbits; some recent examples of literature com-
parisons between detailed dynamical models and TNO orbital distributions are summarized here.
Batygin et al. (2011), Dawson & Murray-Clay (2012), and Morbidelli et al. (2014) all use slightly
different observational constraints to place limits on the exact eccentricity and migration distance of
Neptune's orbit in order to preserve the orbits of cold classical TNOs as observed today. Lawler &
Gladman (2013) test the observed distribution of Kozai Plutinos against the output from a smooth
Neptune migration model (Hahn & Malhotra 2005) and a Nice model simulation (Levison et al.
2008), finding that neither model produces sufficiently high inclinations. Nesvorn´y (2015a) shows
that the timescale of Neptune's migration phase must be fairly slow ((cid:38)10 Myr) in order to replicate
the observed TNO inclination distribution, and Nesvorn´y (2015b) shows that including a "jump" in
Neptune's semimajor axis evolution can create the "kernel" observed in the cold classical TNOs (first
discussed in Petit et al. 2011). Pike et al. (2017) compare the output of a Nice model simulation
(Brasser & Morbidelli 2013) with scattering and resonant TNOs, finding that the population ratios
are consistent with observations except for the 5:1 resonance, which has far more known TNOs than
models would suggest (Pike et al. 2015). Using the observed wide binary TNOs as a constraint on
dynamical evolution suggests that this fragile population formed in-situ (Parker & Kavelaars 2010)
or was emplaced gently into the cold classical region (Fraser et al. 2017). Regardless of the model
involved, using a Survey Simulator is the most accurate and statistically powerful way to make use
of model TNO distributions from dynamical simulations such as these to gain constraints on the
dynamical history of the Solar System.
In this chapter, we discuss what it means for a survey to be "well-characterized" (Section 2) and
explain the structure and function of a Survey Simulator (Section 2.1). In Section 3, we then give
four explicit examples of how to use a Survey Simulator, with actual dynamical model output and
real TNO data. We hope this chapter provides an outline for others to follow.
2. WELL-CHARACTERIZED SURVEYS
A well-characterized survey is one in which the survey field pointings, depths, and tracking fractions
at different magnitudes and on-sky rates of motion have been carefully measured. The largest well-
characterized TNO survey to date is the Outer Solar System Origins Survey (OSSOS; Bannister et al.
2016), which was a large program on the Canada-France-Hawaii Telescope (CFHT) carried out over
five years, and was specifically planned with the Survey Simulator framework in mind. OSSOS builds
on the methodology of three previous well-characterized surveys: the Canada-France Ecliptic Plane
Survey (CFEPS; Jones et al. 2006; Kavelaars et al. 2008b; Petit et al. 2011), the CFEPS high-latitude
component (Petit et al. 2017), and the survey of Alexandersen et al. (2016). The survey design is
discussed extensively in those papers; the OSSOS survey outcomes and parameters in particular are
4
Lawler et al.
presented in detail in Bannister et al. (2016) and Bannister et al. (2018). We here summarize the
main points that are important for creating a well-characterized survey.
These well-characterized surveys are all arranged into observing "blocks:" many individual camera
fields tiled together into a continuous block on the sky (the observing blocks for OSSOS are each
∼10 -- 20 square degrees in size). The full observing block is covered during each dark run (when
the Moon is closest to new) for the 2 months before, 2 months after, and during the dark run
closest to opposition for the observing block. The observing cadence is important for discovering
and tracking TNOs, which change position against the background stars on short timescales. The
time separation between imaging successive camera fields inside each observing block must be long
enough that significant motion against background stars has occurred for TNOs, but not so long that
the TNOs have moved too far to be easily recovered by eye or by software. OSSOS used triplets
of images for each camera field, taken over the course of two hours. The exposure times are chosen
as a careful compromise between photometric depth and limiting trailing1 of these moving sources.
The images are searched for moving objects by an automated and robustly tested moving object
pipeline (CFEPS and OSSOS used the software pipeline described in Petit et al. 2004), and all TNO
candidates discovered by the software are visually inspected. Prior to searching the images, artificial
sources (whose flux and image properties closely mimic the real sample) are inserted into the images.
The detection of these implanted sources marks the fundamental calibration and characterization of
the survey block in photometric depth, detection efficiency, and tracking efficiency. Each observing
block has a measured "filling factor" that accounts for the gaps between CCD chips. Detection and
tracking efficiencies are measured at different on-sky rates of motion using implanted sources in the
survey images, analysed along with the real TNO data. This process is repeated for each observing
block, so each block has known magnitude limits, filling factors, on-sky coverage, and detection and
tracking efficiencies at different on-sky rates of motion. These are then parameterized and become
part of the Survey Simulator.
The tracking of the discovered sample provides another opportunity for biases to enter and the
process must be closely monitored. A survey done in blocks of fields that are repeated ∼monthly
removes the need to make orbit predictions based on only a few hours of arc from a single night's
discovery observations. Such short-arc orbit predictions are notoriously imprecise, and dependence
on them ensures that assumptions made regarding orbit distribution will find their way into the
detected sample as biases. For example, a common assumption for short-arc orbits is a circular orbit.
If follow-up observations based on this circular orbital prediction are attempted with only a small
area of sky coverage, then those orbits whose Keplerian elements match the input assumptions will
be preferentially recovered, while those that do not will be preferentially lost, resulting in a discovery
bias against non-circular orbits. Correcting for this type of ephemeris bias is impossible. Several of
the large-sample TNO surveys had short arcs on a high fraction of the detections; this introduces
unknown tracking biases into the sample that cannot be reproduced in a Survey Simulator because
the systematic reasons for object loss (ephemeris bias, Jones et al. 2010) cannot be modeled as
random. Repeatedly observing the same block of fields, perhaps with some adjustment for the bulk
motion of orbits, helps ensure that ephemeris bias is kept to a minimum. In the OSSOS project we
demonstrate the effectiveness of this approach by managing to track essentially all TNOs brighter
1 We note that clever algorithms can be used to obtain accurate photometry from trailed sources (e.g., Fraser et al.
2016).
OSSOS: X. How to use a Survey Simulator
5
than the flux limits of the discovery sequences (only 2 out of 840 TNOs were not tracked2; Bannister
et al. 2018).
Schwamb et al. (2010) is an example of a large-scale TNO survey from outside our collaboration
that is well-characterized. It has a high tracking fraction and a published pointing history. However,
it has a comparatively noisy and low-resolution detection efficiency function, thus we do not include it
in our Survey Simulator analysis here. Other large sample size TNO surveys have either unpublished
pointings or indications of low tracking fractions leading to unrecoverable ephemeris bias.
A well-characterized survey will have flux limits in each observing block from measurements of
implanted artificial objects, equal sensitivity to a wide range of orbits, and a known spatial coverage
on the sky. A Survey Simulator can now be configured to precisely mimic the observing process for
this survey.
2.1. The Basics of a Survey Simulator
A Survey Simulator allows models of intrinsic Kuiper Belt distributions to be forward-biased to
replicate the biases inherent in a given well-characterized survey. These forward-biased simulated
distributions can then be directly compared with real TNO detections, and a statement can be made
about whether or not a given model is statistically consistent with the known TNOs. One particular
strength of this approach is that the effect of non-detection of certain orbits can be included in the
analysis. Methods that rely on the inversion of orbital distributions are, by their design, not sensitive
to a particular survey's blind spots.
Directly comparing a model with a list of detected TNOs (for example, from the Minor Planet
Center database), with a multitude of unknown detection biases, can lead to inaccurate and possibly
false conclusions. Using a Survey Simulator avoids this problem completely, with the only downside
being that comparisons can only be made using TNOs from well-characterized surveys. Fortunately,
the single OSSOS survey contains over 800 TNOs, and the ensemble of well-characterized affiliated
surveys contains over 1100 TNOs with extremely precisely measured orbits (Bannister et al. 2018).
This is roughly one third of the total number of known TNOs, and one half of the TNOs with orbits
that are well-determined enough to perform dynamical classification.
At its most basic, a Survey Simulator must produce a list of instantaneous on-sky positions, rates
of motion, and apparent magnitudes. These are computed by assigning absolute magnitudes to
simulated objects with a given distribution of orbits. These apparent magnitudes, positions, and
rates of motion are then evaluated to determine the likelihood of detection by the survey, and a
simulated detected distribution of objects is produced. The OSSOS Survey Simulator follows this
basic model, but takes into account more realities of survey limitations. It is the result of refinement
of this Survey Simulator software through several different well-characterized surveys: initially the
CFEPS pre-survey (Jones et al. 2006), then CFEPS (Kavelaars et al. 2008a; Petit et al. 2011), then
Alexandersen et al. (2016), and finally OSSOS (Bannister et al. 2016). The OSSOS Survey Simulator
software and methodology are now robustly tested, and are presented below.
2.2. The Details of the OSSOS Survey Simulator
While the methodology presented here is specific to the OSSOS Survey Simulator, by measuring
on sky pointings, magnitude limits and tracking fractions, a Survey Simulator can be built for any
2 The two objects that were not tracked are d <15 AU Centaurs whose high on-sky rates of motion caused them to
shear off the fields. The possibility of such shearing loss is accounted for in the Survey Simulator.
6
Lawler et al.
survey. The Survey Simulator for the OSSOS ensemble of well-characterized surveys is available
as a package3, and the list of observed characteristics of the TNOs discovered in these surveys is
published in Bannister et al. (2018). To forward-bias a distribution of objects to allow statistical
comparison with the real TNO discoveries in the surveys, the OSSOS Survey Simulator uses the
following steps. The instantaneous on-sky position, rate of motion, and apparent magnitude are
computed from an orbit, position, and H-magnitude, and can be written to a file that contains the
"drawn" simulated objects. This "drawn" file then represents the instantaneous intrinsic distribution
of simulated objects. The Survey Simulator evaluates whether each simulated object falls within one
of the observing blocks of the survey, and if so, uses the tracking and detection efficiency files for that
observing block to calculate whether this object would be detected. If it is detected, the properties of
this object are written to a file containing the simulated detections. A very small fraction of on-sky
motion rates were detected and not tracked in the real survey (Centaurs sometimes shear off the field
due to their high rate of motion), which is accounted for in the Survey Simulator. The simulated
objects are written to the simulated tracked object file with probabilities reflecting this.
The user must supply the Survey Simulator with a routine that generates an object with orbital
elements and an absolute H magnitude. How these are generated is free for the user to decide. The
user may edit the source software to include generation of orbits and absolute magnitudes within
the Survey Simulator, but it is recommended that a separate script is used in conjunction with
the unedited Survey Simulator. The software package comes with a few examples, and details are
provided in the following paragraph showing how we on the OSSOS team have implemented the
generation of simulated TNOs.
The orbital elements of an object can be determined in a variety of ways. The Survey Simulator
can choose an orbit and a random position within that orbit, either from a list of orbits (as would
be produced by a dynamical model) or from a parametric distribution set by the user. Orbits from a
list can also be easily "smeared," that is, variation is allowed within a fraction of the model orbital
elements, in order to smooth a distribution or produce additional similar orbits (however, one must
be careful that the distribution is dominated by the original list of orbits, and not by the specifics
of the smearing procedure). To determine the likely observed magnitude of the source, an absolute
H magnitude is assigned to this simulated object, chosen so as to replicate an H-distribution set
by the user. Tools and examples are provided to set the H-distribution as a single exponential
distribution, to include a knee to a different slope at a given H magnitude (see, for example, Fraser
et al. 2014), or to include a divot in the H-distribution (as in Shankman et al. 2016; Lawler et al.
2018) The H-distribution parameter file also sets the maximum and minimum H values that will be
simulated. The smallest H-magnitude (i.e. largest diameter) is not as important because it is set by
the distribution itself. But the largest H-magnitude (i.e. smallest diameter) is important to match
to the population being simulated. If the maximum H-magnitude is smaller than that of the largest
magnitude TNO in the observational sample one wants to simulate, then the distribution won't be
sensitive to this large H-magnitude tail. If the maximum H value is much larger than that of the
largest magnitude TNO in the sample, it is simply a waste of computational resources, since the
simulation will include many objects that were too faint to be detected in the survey. This does,
3 https://github.com/OSSOS/SurveySimulator
OSSOS: X. How to use a Survey Simulator
7
however, expose a strength of the Survey Simulator, as we can learn from these faint detections that
a hidden reservoir of small objects might exist.
The process of drawing simulated objects and determining if they would have been detected by
the given surveys is repeated until the desired number of simulated tracked or detected objects is
produced by the Survey Simulator. The desired number of simulated detected objects may be the
same as the number of real detected TNOs in a survey in order to measure an intrinsic population size
(as demonstrated in Section 3.3), or an upper limit on a non-detection of a particular subpopulation
(Section 3.4), or it may be a large number in order to test the rejectability of an underlying theoretical
distribution (as demonstrated in Sections 3.1) or to quantify survey biases in a given subpopulation
(Section 3.2).
3. EXAMPLES OF SURVEY SIMULATOR APPLICATIONS
Here we present four examples of different ways to use the Survey Simulator to gain statistically
valuable information about TNO populations. In Section 3.1, we demonstrate how to use the Survey
Simulator to forward-bias the output of dynamical simulations and then statistically compare the
biased simulation with a distribution of real TNOs. In Section 3.2, we use the Survey Simulator to
build a parametric intrinsic distribution and then bias this distribution by our surveys, examining
survey biases for a particular TNO subpopulation in detail. The Survey Simulator can also be used
to measure the size of the intrinsic population required to produce a given number of detections in
a survey; this is demonstrated in Section 3.3. And finally, in Section 3.4, we demonstrate a handy
aspect of the Survey Simulator: using non-detections from a survey to set statistical upper limits on
TNO subpopulations. We hope that these examples will prove useful for dynamical modellers who
want a statistically powerful way to test their models.
3.1. Testing the Output of a Dynamical Model: The Outer Solar System with a Distant Giant
Planet or Rogue Planet
This example expands on the analysis of Lawler et al. (2017), which presents the results of three
different dynamical emplacement simulations of the distant Kuiper Belt. Here we analyze these three
dynamical simulations and also an additional emplacement model that includes a "rogue planet"
that is ejected early in the Solar System's history (Gladman & Chan 2006). The outputs from these
four dynamical TNO emplacement simulations are run through the OSSOS Survey Simulator and
compared with the high pericenter TNOs discovered by OSSOS. The four emplacement simulation
scenarios analyzed here include the following conditions:
1. The four giant planets and including the effects of Galactic tides and stellar flybys (simulation
from Kaib et al. 2011)
2. The four giant planets with an additional 10 Earth mass planet having a = 250 AU and e = 0
(based on a theory proposed in Trujillo & Sheppard 2014), also including the effects of Galactic
tides and stellar flybys (simulation from Lawler et al. 2017).
3. The four giant planets with an additional 10 Earth mass planet having a = 500 AU and e = 0.5
(based on a theory proposed in Batygin & Brown 2016), also including the effects of Galactic
tides and stellar flybys (simulation from Lawler et al. 2017).
4. The four giant planets and an additional 2 Earth mass rogue planet that started with a = 35 AU
and q = 30 AU, which was ejected after ∼200 Myr (simulation from Gladman & Chan 2006).
8
Lawler et al.
The papers which have recently proposed the presence of a distant undiscovered massive planet
(popularly referred to as "Planet 9"; Trujillo & Sheppard 2014; Batygin & Brown 2016) rely on
published detections of large semimajor axis (a), high pericenter distance (q) TNOs, which have
extreme biases against detection in flux-limited surveys4. These large-a, high-q TNOs are drawn from
different surveys, with unpublished and thus unknown observing biases. The Minor Planet Center
(MPC) Database, which provides a repository for TNO detections, does not include information
on the pointings or biases of the surveys that detected these TNOs. Sample selection caused by
survey depth and sky coverage can be non-intuitive, caused by weather, galactic plane avoidance,
and telescope allocation pressure, among other possibilities (Sheppard & Trujillo 2016; Shankman
et al. 2017). Thus the assumption can not be made that a collection of TNOs from different surveys
will be bias-free. Further, the detected sample of objects in the MPC Database provides no insight
into the parts of the sky where a survey may have looked and found nothing, nor to strong variation
in flux sensitivity that occurs due to small variations in image quality.
Here we demonstrate how to use the results of well-characterized surveys to compare real TNO
detections to the output of a dynamical model, comprising a list of orbits. We visualize this output
with a set of cumulative distributions. Figure 1 shows cumulative distributions in six different obser-
vational parameters: semimajor axis a, inclination i, apparent magnitude in r-band mr, pericenter
distance q, distance at detection d, and absolute magnitude in r-band Hr. The outputs from the
four emplacement models are shown by different colours in the plot, with the intrinsic distributions
shown as dotted lines, and the forward-biased simulated detection distributions as solid lines. These
intrinsic distributions have all been cut to only include the pericenter and semimajor axis range pre-
dicted by Batygin & Brown (2016) to be most strongly affected by a distant giant planet: q > 37 AU
and 50 < a < 500 AU.
We cannot directly compare the output from these dynamical simulations covering such a huge range
of a and q to real TNOs because the observing biases present in the detected TNO distributions are
severe. Using a survey simulator combined with a carefully characterized survey, however, allows us
to impose the observing biases onto the simulated sample and determine what the detected simulated
sample would have been. We use the OSSOS Survey Simulator and the q > 37 AU, 50 < a < 500 AU
TNOs discovered by the OSSOS ensemble in this example (Figure 1).
The solid lines in Figure 1 show the forward-biased simulated detections from the Survey Simulator,
and the black points show the real TNOs detected in the OSSOS ensemble with the same a and q cut.
These can now be directly compared, as the same biases have been applied. The effect of observing
biases varies widely among the six parameters plotted, and can be seen by comparing the intrinsic
simulated distributions (dotted lines) with the corresponding biased simulated distributions (solid
lines) in Figure 1. Unsurprisingly, these surveys are biased toward detecting the smallest-a, lower
inclination, and lowest-q TNOs.
These models have not been explicitly constructed in an attempt to produce the orbital and magni-
tude distribution of the elements in the detection range, so the following exercise is pedagogic rather
than attempting to be diagnostic. All these models statistically fail dramatically in producing several
of the distributions shown here (discussed in detail later in this section), but how they fail allows
one to understand what changes to the model may be required. All the models generically produce
4 Brown (2017) does attempt a numerical simulation to calculate likelihood of detection for different orbital param-
eters in the high-q population, but this is backing out the underlying distribution in the underlying sample and still
relies on assumptions about the unknown sensitivity and completeness of the surveys.
OSSOS: X. How to use a Survey Simulator
9
Figure 1. Cumulative distributions of TNOs in six different parameters: semimajor axis a, inclination
i, apparent r-band magnitude mr, pericenter distance q, distance at detection d, and absolute magnitude
in r-band Hr. The result of four different emplacement models are presented: the known Solar System
(orange), the Solar System plus a circular orbit Planet 9 (blue), or plus an eccentric orbit Planet 9 (gray),
and a "rogue" planet simulation (purple). The intrinsic model distributions are shown with dotted lines,
and the resulting simulated detections in solid lines. Black circles show non-resonant TNOs discovered by
the OSSOS ensemble having q > 37 AU and 50 < a < 500 AU.
a slight misbalance in both the absolute (lower right panel of Figure 1) and apparent magnitude dis-
tributions (upper right panel of Figure 1), with too few bright objects being present in the predicted
sample, when using the input Hr-magnitude distribution from Lawler et al. (2018). Changes in the
H-magnitude exponents or break points that are comparable to their current uncertainties will be
sufficient to produce a greatly improved match, so this comparison is less interesting than for the
orbital distributions.
The orbital inclination distribution (upper center panel of Figure 1) of the detected sample (black
dots) is roughly uniform up to about i = 18◦, it has few members from 18−25◦, and then has the final
one third of the sample distributed up to about 40◦. This observed cutoff near 40◦ is strongly affected
by survey biases, as evidenced by the dramatic elimination of this large fraction of the intrinsic model
population that is present in the two Planet 9 simulations (compare the distribution of dotted and
Lawler et al.
10
solid blue and gray lines). In contrast, the relative dearth in the 18− 25◦ range of the observed TNOs
is not caused by the survey biases: none of the biased models show this effect. The rogue planet
model's intrinsic distribution (dotted purple line) is a relatively good match to the detections up to
about i = 15◦; when biased by the Survey Simulator (solid purple line) this model predicts far too
cold a distribution. The rogue model used here was from a simulation with a nearly-coplanar initial
rogue planet; simulations with an initially inclined extra planet will give higher inclinations for the
Kuiper Belt objects and would thus be required in this scenario. The two Planet 9 scenarios shown
give better (but still rejectable) comparisons to the detections.
The comparison of the models in the semimajor axis distribution gives very clear trends (upper left
panel of Figure 1). The rogue planet model (solid purple line) used seriously underpredicts the fraction
of large-a TNOs but does produce the abundant a < 70 AU objects in the detected sample (black
dots). This is caused by the rogue spending little time in the enormous volume beyond 100 AU and
thus being unable to significanly lift perihelia for those semimajor axes. In contrast, the control (solid
orange line) and Planet 9 models (solid blue and gray lines) greatly underpredict the a < 100 AU
fraction because the distant planet has very little dynamical effect on these relatively tightly-bound
orbits; these models predict roughly the correct fraction, ∼20% of objects having a > 100 AU, but
the distribution of larger-a TNOs is more extended in reality than these models predict. Lastly, the
q > 37 AU perihelion distribution (lower left panel of Figure 1) of the rogue model (solid purple line)
is more extended than the real objects, due to this particular rogue efficiently raising the perihelia
of the a = 50 − 100 AU populations by a few AU, to a broad distribution. The control model's q
distribution (solid orange line) is far too concentrated, while the Planet 9 simulations (solid blue and
gray lines) qualitatively provide the match the real TNOs in this distribution.
In order to determine whether any of these biased distributions provide a statistically acceptable
match to the real TNOs, we use a bootstrapped Anderson-Darling statistic5 (Anderson & Darling
1954), calculated for each of the six distributions. This technique is described in detail in previ-
ous literature (e.g. Kavelaars et al. 2009; Petit et al. 2011; Bannister et al. 2016; Shankman et al.
2016), and we summarize below. An Anderson-Darling statistic is calculated for the simulated bi-
ased distributions compared with the real TNOs; this statistic is summed over all distributions being
tested. The Anderson-Darling statistic is then bootstrapped by drawing a handful of random sim-
ulated objects from each simulated distribution, calculating the resulting AD statistics, summing
over all distributions, then comparing this to how often the summed AD statistic for the real TNOs
occurs (following Parker 2015; Alexandersen et al. 2016). If a summed AD statistic as large as the
summed AD statistic for the real TNOs occurs in < 5% of randomly drawn samples, we conclude this
simulated distribution is inconsistent with observations and we can reject it at the 95% confidence
level.
When we calculate the bootstrapped AD statistics for each of the simulated distributions as com-
pared with the real TNOs in Figure 1, we find that all four of the tested simulations are inconsistent
with the data and we reject all of them at > 99% confidence level. This may be surprising to
those unaccustomed to these comparisons. Matches between data and models that are not statisti-
cally rejectable have almost no noticeable differences between the data and the biased model in all
parameters than are compared (see, e.g., Figure 3 in Gladman et al. 2012).
5 The Anderson-Darling test is similar to the better-known Kolmogorov-Smirnov test, but with higher sensitivity to
the tails of the distributions being compared.
OSSOS: X. How to use a Survey Simulator
11
We note that none of the four dynamical emplacement models analysed here include the effects
of Neptune's migration, which is well-known to have an important influence on the structure of the
distant Kuiper belt. Recent detailed migration simulations (Kaib & Sheppard 2016; Nesvorn´y et al.
2016; Pike et al. 2017; Pike & Lawler 2017) have shown that temporary resonance capture and Kozai
cycling within resonances during Neptune's orbital evolution has important effects on the overall
distribution of distant TNOs, particularly in raising pericenters and semimajor axes. Incorporating
the effects of Neptune's migration may produce a better fit between the real TNOs and the models
shown in Figure 1.
We reiterate that the point of this section has been to provide a walk-through of how to compare the
output of a dynamical model to real TNO detections in a statistically powerful way. The preceding
discussion of the shortcomings of the specific dynamical models presented here highlights that a
holistic approach to dynamical simulations of Kuiper belt emplacement is necessary. For example,
Nesvorn´y (2015a) uses the CFEPS Survey Simulator and CFEPS-discovered TNOs to constrain a
Neptune migration model presented in that work, and Shankman et al. (2016) uses the OSSOS Survey
Simulator and TNOs to improve the dynamical emplacement model of Kaib et al. 2011. We hope
that use of a Survey Simulator will become standard practice for testing dynamical emplacement
models in the future.
3.2. Using a Parametric Population Distribution: Biases in Detection of Resonant TNOs
In this section we show how to use the Survey Simulator with a resonant TNO population in
order to demonstrate two important and useful points: how to build a simulated population from
a parametric distribution, and showing how the Survey Simulator handles longitude biases in non-
uniform populations. For this demonstration, we build a toy model of the 2:1 mean motion resonance
with Neptune, using a parametric distribution built within the Survey Simulator software (though we
note that this parametric distribution could just as easily be created with a separate script and later
utilized by the unedited Survey Simulator). The parametric distribution here is roughly based on
that used in Gladman et al. (2012), but greatly simplified and not attempting to match the real 2:1
TNO distributions. The Survey Simulator chooses from this parametric distribution with a within
±0.5 AU of the resonance center (47.8 AU), q = 30 AU, i = 0◦,6 and libration amplitudes 0 − 10◦.
There are three resonant islands in the 2:1 resonance, and the libration center (cid:104)φ21(cid:105) is chosen to
populate these three islands equally in this toy model. For the purposes of our toy model, we define
these angles as: leading asymmetric (cid:104)φ21(cid:105) = 80◦, symmetric (cid:104)φ21(cid:105) = 180◦, and trailing asymmetric
(cid:104)φ21(cid:105) = 280◦. The resonant angle φ21 in the snapshot is chosen sinusoidally within the libration
amplitude. The ascending node Ω and mean anomaly M are chosen randomly, then the argument
of pericenter ω is chosen to satisfy the resonant condition φ21 = 2λTNO − λN − TNO, where mean
longitude λ = Ω + ω + M, longitude of pericenter = Ω + ω, and the subscripts TNO and N
denote the orbital elements of the TNO and Neptune, respectively. The last step is to assign an H
magnitude, in this case from a literature TNO H-distribution (Lawler et al. 2018). As each simulated
TNO is drawn from this distribution, its magnitude (resulting from its instantaneous distance and
H magnitude), on-sky position, and rate of on-sky motion are evaluated by the Survey Simulator to
ascertain whether or not this TNO would have been detected by the survey for which configurations
are provided to the simulator.
6 The inclinations are actually set to a very small value close to zero to avoid ambiguities in the other orbital angles.
12
Lawler et al.
Figure 2. Black points show the positions of TNOs in a snapshot from this parametric toy model of TNOs
in the 2:1 mean-motion resonance. The position of Neptune is shown by a blue circle, and dotted circles
show distances from the Sun. Red points show simulated detections after running this model through the
Survey Simulator. One third of the TNOs are in each libration island in the intrinsic model, but 2/3 of the
detections are in the leading asymmetric island (having pericenters in the upper right quadrant of this plot).
This bias is simply due to the longitudinal direction of pointings within the OSSOS ensemble and pericenter
locations in this toy model of the 2:1 resonance.
Figure 2 shows the results of this simulation. Due to the (unrealistically) low libration amplitudes
in this toy model simulation, the three resonant islands show up as discrete sets of orbits, with
pericenters in three clusters: symmetric librators are opposite Neptune, and leading and trailing are
in the upper right and lower left of the plot, respectively7. While the intrinsic distributions in this
7 We note that in reality the symmetric island generally has very large libration amplitudes and the pericenter
positions of symmetric librators overlap with each of the asymmetric islands (e.g. Volk et al. 2016), thus diagnosing
the membership of each island is not as simple as this toy model makes it appear.
OSSOS: X. How to use a Survey Simulator
13
model are evenly distributed among the three islands by design (33% in each), the detected TNOs
are not. This is because TNOs on eccentric orbits are most likely to be detected close to pericenter,
and in this toy model, the pericenters are highly clustered around three distinct positions on the
sky.8
The red points in Figure 2 show the simulated TNOs that are detected by the OSSOS Survey
Simulator. Just because of the positions of survey pointings on the sky, the detections are highly
biased toward the leading island, which has over 2/3 of the detections, with fewer detections in the
trailing island, and only 10% of all the detections in the symmetric island. Without knowing the
on-sky detection biases (as is the case for TNOs pulled from the MPC database), one could easily
(but erroneously) conclude that the leading island is much more populated than the trailing island,
when in fact the intrinsic populations are equal.
The relative fraction of n:1 TNOs that inhabit different resonant islands has been discussed in the
literature as a diagnostic of Neptune's past migration history (Chiang & Jordan 2002; Murray-Clay &
Chiang 2005; Pike & Lawler 2017). Past observational studies have relied on n:1 resonators from the
MPC database with unknown biases, or the handful of n:1 resonators detected in well-characterized
surveys to gain weak (but statistically tested) constraints on the relative populations. Upcoming
detailed analysis using the full OSSOS survey and making use of the Survey Simulator will provide
stronger constraints on this fraction, without worries about unknown observational biases (Chen,
Y.-T., private communication).
3.3. Determining Intrinsic Population Size: The Centaurs
The Survey Simulator can easily be used to determine intrinsic population sizes. As described in
Section 2.1, the Survey Simulator keeps track of the number of "drawn" simulated objects that are
needed for the requested number of simulated tracked objects. By asking the Survey Simulator to
produce the same number of tracked simulated objects as were tracked in a given survey, the number
of drawn simulated objects is a realization of the intrinsic population required for the survey to have
detected the actual number of TNOs that were found by the survey. By repeating this many times,
with different random number seeds, different orbits and instantaneous positions are chosen from the
model and slightly different numbers of simulated drawn objects are required each time. This allows
us to measure the range of intrinsic population sizes needed to produce a given number of tracked,
detected TNOs in a survey.
Here we measure the intrinsic population required to produce the 17 Centaurs that are detected in
the OSSOS ensemble. Once the parameters of the orbital distribution and the H-distribution have
been pinned down using the AD statistical analysis outlined in Section 3.1 (this is done in detail
for the Centaurs in Lawler et al. 2018), we run the Survey Simulator until it produces 17 tracked
Centaurs from the a < 30 AU portion of the Kaib et al. (2011) scattering TNO model, and record
the number of simulated drawn objects required. As the OSSOS ensemble discovered Centaurs down
to Hr (cid:39) 14, the Survey Simulator is run to this H-magnitude limit. We repeat this 1000 times to
find the range of intrinsic population sizes that can produce 17 simulated tracked objects.
Using the properties of simulated objects in the drawn file, we measure the intrinsic population
size to Hr < 12 (right panel, Figure 3). The median intrinsic population required for 17 detections is
8 While this is obviously an exaggerated example, pericenters in all of the resonant TNO populations are most
likely to occur at certain sky positions, and thus this is an important consideration when discussing survey biases (see
Gladman et al. 2012; Lawler & Gladman 2013; Volk et al. 2016, for detailed discussions and analysis of these effects
for resonant populations).
14
Lawler et al.
Figure 3. The range of intrinsic Centaur population sizes required for the Survey Simulator to produce the
same number of Centaur detections (17) as were discovered by the OSSOS ensemble, Hr < 8.66 intrinsic
populations shown in red (left panel) and Hr < 12 shown in orange (right panel). The solid lines highlight the
median population, and dotted lines show the 95% upper and lower confidence limits on the intrinsic Centaur
population for each H-magnitude limit. Using the OSSOS ensemble, we measure an intrinsic populations of
130+80−60 Hr < 8.66 Centaurs and 3700+2100−1600 Hr < 12 Centaurs with 95% confidence.
3700, with 97.5% of population estimates falling above 2100, and 97.5% falling below 5800 (these two
values bracket 95% of the population estimates). The result is a statistically tested 95% confidence
limit on the intrinsic Centaur population of 3700+2100−1600 for Hr < 12.
To ease comparison with other statistically produced intrinsic TNO population estimates (e.g.
Petit et al. 2011; Gladman et al. 2012), we also measure a population estimate for Hr < 8.66, which
corresponds to D (cid:38) 100 km9 (left panel, Figure 3). Even though none of the detected Centaurs in
OSSOS had Hr magnitudes this small, this estimate is valid because it is calculated from our Survey
Simulator-based population estimate and a measured H-magnitude distribution (from Lawler et al.
2018). For Hr < 8.66, the statistically tested 95% confidence limit on the intrinsic Centaur population
is 130+80−60.
3.4. Constraints from Non-Detections: Testing a Theoretical Distant Population
Here we show a perhaps unintuitive aspect of the Survey Simulator: non-detections can be just as
powerful as detections for constraining Kuiper Belt populations. Non-detections can only be used if
the full pointing list from a survey is published along with the detected TNOs. An examination of
the orbital distribution of TNOs in the MPC database makes it clear that there is a sharp dropoff
in the density of TNO detections at a (cid:38) 50 AU and q (cid:38) 40 AU (Figure 2 in Sheppard & Trujillo
2016, demonstrates this beautifully). Is this the result of observation bias or a real dropoff? Without
9 The approximate Hr magnitude that corresponds to D of 100 km is calculated assuming an albedo of 0.04 and
using an average plutino colour g − r = 0.5 (Alexandersen et al. 2016).
OSSOS: X. How to use a Survey Simulator
15
carefully accounting for survey biases (Allen et al. 2001, 2002), by using the MPC database alone,
there is no way to know.
Figure 4. Coloured points show the relative populations and semimajor axis-eccentricity distributions
from the CFEPS L7 model of the Kuiper belt, where absolute population estimates have been produced
for each subpopulation in the well-characterized CFEPS survey (Petit et al. 2011; Gladman et al. 2012),
populations are scaled to Hg < 8.5 and colour-coded by dynamical class. Resonances included in this model
(those with >1 TNO detected by CFEPS) are labelled. A low-e artificial test population has been injected
at 60 AU (black points); the number of points shows the upper limit on this population determined by zero
detections in OSSOS, extrapolated to match the CFEPS model (Hr < 8).
Randomly drawing zero objects when you expect three has a probability of 5%, assuming Poisson
statistics; thus, the simulated population required to produce three detections is the 95% confidence
upper limit for a population that produced zero detections. As our example for non-detection upper
limits, we create an artificial population in the distant, low-eccentricity Kuiper belt, where OSSOS
has zero detections. Figure 4 shows the CFEPS L7 debiased Kuiper belt10 (Petit et al. 2011; Gladman
et al. 2012), where this low-eccentricity artificial population has been inserted at 60 AU (small black
points) to demonstrate the power of non-detections, even in parameter space with low sensitivity
10 Model available at www.cfeps.net
16
Lawler et al.
(i.e., low-e TNOs at 60 AU will never come very close to the Sun, thus will always remain faint and
difficult to detect). We run this artificial population through the Survey Simulator until we have
three detections in order to measure a 95% confidence upper limit on this population. At this large
distance, we are only sensitive to relatively large TNOs (Hr (cid:46) 7, corresponding to D (cid:38) 180 km for
an albedo of 0.04). For there to be an expectation to detect three objects in the survey fields, the
Survey Simulator tells us that there would need to exist a population of ∼90 such objects, which
is our upper limit on this population. Extrapolating this population to match the H magnitude
limits of the CFEPS L7 subpopulations plotted in Figure 4, the total number of which are scaled
appropriately for Hg < 8.5, gives ∼700 objects in this artificial population with Hr < 8.
This is a very small population size. We note that this specific analysis is only applicable to
dynamically cold TNOs of relatively large size (Hr < 7). A steep size distribution could allow
many smaller TNOs to remain undiscovered on similar orbits. The point of this exercise it to show
statistically tested constraints on populations with no survey detections. For comparison, at this
absolute magnitude limit (Hr < 7), the scattering disk is estimated to have an intrinsic population
of ∼4000 (Lawler et al. 2018), the plutinos are estimated to have an intrinsic population of ∼500
(Volk et al. 2016), and the detached TNOs are estimated to number ∼4000 (Petit et al. 2011). Using
the estimated populations and size distribution slopes from Petit et al. (2011) gives 3000 Hr < 7
TNOs in the classical belt. The 3:1 mean-motion resonance, which is located at a similar semimajor
axis (a = 62.6 AU), is estimated to have an Hr < 7 population of ∼200 (Gladman et al. 2012). The
3:1 TNOs, however, are much more easily detected in a survey due to their higher eccentricities as
compared with our artificial 60 AU population, thus a given survey would be sensitive to a larger
H-magnitude for the 3:1 population than the 60 AU cold test population. This statistically tested
population limit essentially shows that not very many low eccentricity, distant TNOs can be hiding
from the OSSOS survey ensemble, especially when compared with other TNO populations.
4. CONCLUSION
Using TNO discoveries from well-characterized surveys and only analysing the goodness of fit
between models and TNO discoveries after forward-biasing the models gives a statistically powerful
framework within which to validate dynamical models of Kuiper belt formation. Understanding the
effects that various aspects of Neptune's migration have on the detailed structure of the Kuiper belt
not only provides constraints on the formation of Neptune and Kuiper belt planetesimals, but also
provides useful comparison to extrasolar planetesimal belts (Matthews & Kavelaars 2016). There
are, of course, many lingering mysteries about the structure of the Kuiper belt, some of which may
be solved by detailed dynamical simulations in combination with new TNO discoveries in the near
future.
One such mystery is explaining the very large inclinations in the scattering disk (Shankman et al.
2016; Lawler et al. 2018) and inside mean-motion resonances (Gladman et al. 2012). Some possible
dynamical mechanisms to raise inclinations include rogue planets/large mass TNOs (Gladman &
Chan 2006; Silsbee & Tremaine 2017), interactions with a distant massive planet (Gomes et al. 2015;
Lawler et al. 2017), and diffusion from the Oort cloud (Kaib et al. 2009; Brasser et al. 2012). These
theories may also be related to the recent discovery that the mean plane of the distant Kuiper belt
is warped (Volk & Malhotra 2017).
Explaining the population of high pericenter TNOs (Sheppard & Trujillo 2016; Shankman et al.
2017) is also difficult with current Neptune migration models. Theories to emplace high pericenter
OSSOS: X. How to use a Survey Simulator
17
TNOs include dynamical diffusion (Bannister et al. 2017), dropouts from mean-motion resonances
during grainy Neptune migration (Kaib & Sheppard 2016; Nesvorn´y et al. 2016), dropouts from mean-
motion resonances during Neptune's orbital circularization phase (Pike & Lawler 2017), interactions
with a distant giant planet (Gomes et al. 2015; Batygin & Brown 2016; Lawler et al. 2017), a stellar
flyby (Morbidelli & Levison 2004), capture from a passing star (J´ılkov´a et al. 2015), and perturbations
in the Solar birth cluster (Brasser & Schwamb 2015).
Here we have highlighted only a small number of the inconsistencies between models and real TNO
orbital data. The level of detail that must be included in Neptune migration simulations has increased
dramatically with the release of the full OSSOS dataset, containing hundreds of TNOs with the most
precise orbits ever measured. The use of the Survey Simulator will be vital for testing future highly
detailed dynamical emplacement simulations, and for solving the lingering mysteries in the observed
structure of the Kuiper belt.
SML gratefully acknowledges support from the NRC-Canada Plaskett Fellowship. MTB is sup-
ported by UK STFC grant ST/L000709/1.
The authors acknowledge the sacred nature of Maunakea, and appreciate the opportunity to observe
from the mountain. CFHT is operated by the National Research Council (NRC) of Canada, the
Institute National des Sciences de l'Universe of the Centre National de la Recherche Scientifique
(CNRS) of France, and the University of Hawaii, with OSSOS receiving additional access due to
contributions from the Institute of Astronomy and Astrophysics, Academia Sinica, Taiwan. Data
are produced and hosted at the Canadian Astronomy Data Centre; processing and analysis were
performed using computing and storage capacity provided by the Canadian Advanced Network For
Astronomy Research (CANFAR).
REFERENCES
Adams, F. C. 2010, ARA&A, 48, 47,
Batygin, K., & Brown, M. E. 2016, AJ, 151, 22,
doi: 10.1146/annurev-astro-081309-130830
doi: 10.3847/0004-6256/151/2/22
Alexandersen, M., Gladman, B., Kavelaars, J. J.,
Batygin, K., Brown, M. E., & Betts, H. 2012,
et al. 2016, AJ, 152, 111,
doi: 10.3847/0004-6256/152/5/111
Allen, R. L., Bernstein, G. M., & Malhotra, R.
2001, ApJL, 549, L241, doi: 10.1086/319165
-- . 2002, AJ, 124, 2949, doi: 10.1086/343773
Anderson, T. W., & Darling, D. A. 1954, Journal
of the American Statistical Association, 49, 765,
doi: 10.1080/01621459.1954.10501232
Bannister, M. T., Kavelaars, J. J., Petit, J.-M.,
et al. 2016, AJ, 152, 70,
doi: 10.3847/0004-6256/152/3/70
Bannister, M. T., Shankman, C., Volk, K., et al.
2017, AJ, 153, 262,
doi: 10.3847/1538-3881/aa6db5
Bannister, M. T., Gladman, B. J., Kavelaars,
J. J., et al. 2018, AJ, 152, 70,
doi: 10.3847/0004-6256/152/3/70
ApJL, 744, L3,
doi: 10.1088/2041-8205/744/1/L3
Batygin, K., Brown, M. E., & Fraser, W. C. 2011,
ApJ, 738, 13, doi: 10.1088/0004-637X/738/1/13
Bernstein, G. M., Trilling, D. E., Allen, R. L.,
et al. 2004, AJ, 128, 1364, doi: 10.1086/422919
Brasser, R., Duncan, M. J., Levison, H. F.,
Schwamb, M. E., & Brown, M. E. 2012, Icarus,
217, 1, doi: 10.1016/j.icarus.2011.10.012
Brasser, R., & Morbidelli, A. 2013, Icarus, 225,
40, doi: 10.1016/j.icarus.2013.03.012
Brasser, R., & Schwamb, M. E. 2015, MNRAS,
446, 3788, doi: 10.1093/mnras/stu2374
Brown, M. E. 2017, AJ, 154, 65,
doi: 10.3847/1538-3881/aa79f4
Brunini, A., & Fernandez, J. A. 1996, A&A, 308,
988
18
Lawler et al.
Chiang, E. I., & Jordan, A. B. 2002, AJ, 124,
Jones, R. L., Gladman, B., Petit, J.-M., et al.
3430, doi: 10.1086/344605
Davis, D. R., & Farinella, P. 1997, Icarus, 125, 50,
doi: 10.1006/icar.1996.5595
2006, Icarus, 185, 508,
doi: 10.1016/j.icarus.2006.07.024
Kaib, N. A., Roskar, R., & Quinn, T. 2011, Icarus,
Dawson, R. I., & Murray-Clay, R. 2012, ApJ, 750,
215, 491, doi: 10.1016/j.icarus.2011.07.037
43, doi: 10.1088/0004-637X/750/1/43
Duncan, M., Quinn, T., & Tremaine, S. 1988,
ApJL, 328, L69, doi: 10.1086/185162
Fernandez, J. A. 1980, MNRAS, 192, 481,
doi: 10.1093/mnras/192.3.481
Fraser, W., Alexandersen, M., Schwamb, M. E.,
et al. 2016, AJ, 151, 158,
doi: 10.3847/0004-6256/151/6/158
Fraser, W. C., Brown, M. E., Morbidelli, A.,
Parker, A., & Batygin, K. 2014, ApJ, 782, 100,
doi: 10.1088/0004-637X/782/2/100
Fraser, W. C., & Kavelaars, J. J. 2009, AJ, 137,
72, doi: 10.1088/0004-6256/137/1/72
Fraser, W. C., Bannister, M. T., Pike, R. E., et al.
2017, Nature Astronomy, 1, 0088,
doi: 10.1038/s41550-017-0088
Gladman, B., & Chan, C. 2006, ApJL, 643, L135,
doi: 10.1086/505214
Gladman, B., Kavelaars, J. J., Nicholson, P. D.,
Loredo, T. J., & Burns, J. A. 1998, AJ, 116,
2042, doi: 10.1086/300573
Gladman, B., Lawler, S. M., Petit, J.-M., et al.
2012, AJ, 144, 23,
doi: 10.1088/0004-6256/144/1/23
Gomes, R., Levison, H. F., Tsiganis, K., &
Morbidelli, A. 2005, Nature, 435, 466,
doi: 10.1038/nature03676
Gomes, R. S., Soares, J. S., & Brasser, R. 2015,
Icarus, 258, 37, doi: 10.1016/j.icarus.2015.06.020
Hahn, J. M., & Malhotra, R. 2005, AJ, 130, 2392,
doi: 10.1086/452638
Hamid, S. E., Marsden, B. G., & Whipple, F. L.
1968, AJ, 73, 727, doi: 10.1086/110685
Irwin, M., Tremaine, S., & Zytkow, A. N. 1995,
AJ, 110, 3082, doi: 10.1086/117749
Jewitt, D., & Luu, J. 1993, Nature, 362, 730,
doi: 10.1038/362730a0
Jewitt, D. C., & Luu, J. X. 1995, AJ, 109, 1867,
doi: 10.1086/117413
J´ılkov´a, L., Portegies Zwart, S., Pijloo, T., &
Hammer, M. 2015, MNRAS, 453, 3157,
doi: 10.1093/mnras/stv1803
Jones, R. L., Parker, J. W., Bieryla, A., et al.
2010, AJ, 139, 2249,
doi: 10.1088/0004-6256/139/6/2249
Kaib, N. A., & Sheppard, S. S. 2016, AJ, 152, 133,
doi: 10.3847/0004-6256/152/5/133
Kaib, N. A., Becker, A. C., Jones, R. L., et al.
2009, ApJ, 695, 268,
doi: 10.1088/0004-637X/695/1/268
Kavelaars, J., Jones, L., Gladman, B., Parker,
J. W., & Petit, J.-M. 2008a, The Orbital and
Spatial Distribution of the Kuiper Belt, ed.
M. A. Barucci, H. Boehnhardt, D. P.
Cruikshank, A. Morbidelli, & R. Dotson
(University of Arizona Press, Tucson), 59 -- 69
Kavelaars, J. J., Gladman, B., Petit, J., Parker,
J. W., & Jones, L. 2008b, in Bulletin of the
American Astronomical Society, Vol. 40,
AAS/Division for Planetary Sciences Meeting
Abstracts #40, 481
Kavelaars, J. J., Jones, R. L., Gladman, B. J.,
et al. 2009, AJ, 137, 4917,
doi: 10.1088/0004-6256/137/6/4917
Kobayashi, H., & Ida, S. 2001, Icarus, 153, 416,
doi: 10.1006/icar.2001.6700
Lawler, S. M., & Gladman, B. 2013, AJ, 146, 6,
doi: 10.1088/0004-6256/146/1/6
Lawler, S. M., Shankman, C., Kaib, N., et al.
2017, AJ, 153, 33,
doi: 10.3847/1538-3881/153/1/33
Lawler, S. M., Shankman, C., Kavelaars, J. J.,
et al. 2018, AJ, 155, 197,
doi: 10.3847/1538-3881/aab8ff
Levison, H. F., Morbidelli, A., Vanlaerhoven, C.,
Gomes, R., & Tsiganis, K. 2008, Icarus, 196,
258, doi: 10.1016/j.icarus.2007.11.035
Lykawka, P. S., & Mukai, T. 2008, AJ, 135, 1161,
doi: 10.1088/0004-6256/135/4/1161
Malhotra, R. 1993, Nature, 365, 819,
doi: 10.1038/365819a0
-- . 1995, AJ, 110, 420, doi: 10.1086/117532
Matthews, B. C., & Kavelaars, J. 2016, SSRv,
205, 213, doi: 10.1007/s11214-016-0249-0
Morbidelli, A., Gaspar, H. S., & Nesvorny, D.
2014, Icarus, 232, 81,
doi: 10.1016/j.icarus.2013.12.023
Morbidelli, A., & Levison, H. F. 2004, AJ, 128,
2564, doi: 10.1086/424617
OSSOS: X. How to use a Survey Simulator
19
Murray-Clay, R. A., & Chiang, E. I. 2005, ApJ,
619, 623, doi: 10.1086/426425
Nesvorn´y, D. 2015a, AJ, 150, 73,
doi: 10.1088/0004-6256/150/3/73
-- . 2015b, AJ, 150, 68,
doi: 10.1088/0004-6256/150/3/68
Nesvorn´y, D., & Vokrouhlick´y, D. 2016, ApJ, 825,
94, doi: 10.3847/0004-637X/825/2/94
Nesvorn´y, D., Vokrouhlick´y, D., & Morbidelli, A.
2013, ApJ, 768, 45,
doi: 10.1088/0004-637X/768/1/45
Nesvorn´y, D., Vokrouhlick´y, D., & Roig, F. 2016,
ApJL, 827, L35,
doi: 10.3847/2041-8205/827/2/L35
Parker, A. H. 2015, Icarus, 247, 112,
doi: 10.1016/j.icarus.2014.09.043
Parker, A. H., & Kavelaars, J. J. 2010, ApJL, 722,
L204, doi: 10.1088/2041-8205/722/2/L204
Petit, J., Kavelaars, J. J., Gladman, B., et al.
2011, AJ
Petit, J.-M., Holman, M., Scholl, H., Kavelaars,
J., & Gladman, B. 2004, MNRAS, 347, 471,
doi: 10.1111/j.1365-2966.2004.07217.x
Petit, J.-M., Morbidelli, A., & Valsecchi, G. B.
1999, Icarus, 141, 367,
doi: 10.1006/icar.1999.6166
Petit, J.-M., Kavelaars, J. J., Gladman, B. J.,
et al. 2017, AJ, 153, 236,
doi: 10.3847/1538-3881/aa6aa5
Pike, R. E., Kavelaars, J. J., Petit, J. M., et al.
2015, AJ, 149, 202,
doi: 10.1088/0004-6256/149/6/202
Pike, R. E., & Lawler, S. 2017, AJ, 153, 127,
doi: 10.3847/1538-3881/aa5be9
Pike, R. E., Lawler, S., Brasser, R., et al. 2017,
AJ, 153, 127, doi: 10.3847/1538-3881/aa5be9
Schwamb, M. E., Brown, M. E., & Rabinowitz,
D. L. 2009, ApJL, 694, L45,
doi: 10.1088/0004-637X/694/1/L45
Schwamb, M. E., Brown, M. E., Rabinowitz,
D. L., & Ragozzine, D. 2010, ApJ, 720, 1691,
doi: 10.1088/0004-637X/720/2/1691
Shankman, C., Kavelaars, J., Gladman, B. J.,
et al. 2016, AJ, 151, 31,
doi: 10.3847/0004-6256/151/2/31
Shankman, C., Kavelaars, J. J., Bannister, M. T.,
et al. 2017, AJ, 154, 50,
doi: 10.3847/1538-3881/aa7aed
Sheppard, S. S., & Trujillo, C. 2016, AJ, 152, 221,
doi: 10.3847/1538-3881/152/6/221
Silsbee, K., & Tremaine, S. 2017, ArXiv e-prints.
https://arxiv.org/abs/1712.03961
Stern, S. A. 1995, AJ, 110, 856,
doi: 10.1086/117568
-- . 1996, A&A, 310, 999
Thommes, E. W., Duncan, M. J., & Levison, H. F.
1999, Nature, 402, 635, doi: 10.1038/45185
Trujillo, C. A., Jewitt, D. C., & Luu, J. X. 2001,
AJ, 122, 457, doi: 10.1086/321117
Trujillo, C. A., & Sheppard, S. S. 2014, Nature,
507, 471, doi: 10.1038/nature13156
Tsiganis, K., Gomes, R., Morbidelli, A., &
Levison, H. F. 2005, Nature, 435, 459,
doi: 10.1038/nature03539
Volk, K., & Malhotra, R. 2017, AJ, 154, 62,
doi: 10.3847/1538-3881/aa79ff
Volk, K., Murray-Clay, R., Gladman, B., et al.
2016, AJ, 152, 23,
doi: 10.3847/0004-6256/152/1/23
|
1207.0836 | 1 | 1207 | 2012-07-03T21:10:51 | Searching for the first Near-Earth Object family | [
"astro-ph.EP"
] | We report on our search for genetically related asteroids amongst the near-Earth object (NEO) population - families of NEOs akin to the well known main belt asteroid families. We used the technique proposed by Fu et al. (2005) supplemented with a detailed analysis of the statistical significance of the detected clusters. Their significance was assessed by comparison to identical searches performed on 1,000 'fuzzy-real' NEO orbit distribution models that we developed for this purpose. The family-free 'fuzzy-real' NEO models maintain both the micro and macro distribution of 5 orbital elements (ignoring the mean anomaly). Three clusters were identified that contain four or more NEOs but none of them are statistically significant at \geq 3{\sigma}. The most statistically significant cluster at the \sim 2{\sigma} level contains 4 objects with H < 20 and all members have long observational arcs and concomitant good orbital elements. Despite the low statistical significance we performed several other tests on the cluster to determine if it is likely a genetic family. The tests included examining the cluster's taxonomy, size-frequency distribution, consistency with a family-forming event during tidal disruption in a close approach to Mars, and whether it is detectable in a proper element cluster search. None of these tests exclude the possibility that the cluster is a family but neither do they confirm the hypothesis. We conclude that we have not identified any NEO families. | astro-ph.EP | astro-ph |
Searching for the first Near-Earth Object family
Eva Schunov´aa, Mikael Granvikb,e, Robert Jedickeb, Giovanni Gronchic, Richard
Wainscoatb, Shinsuke Abed
E-mail: [email protected]
aDepartment of Astronomy, Physics of the Earth and Meteorology, Comenius University,
Mlynsk´a dolina, Bratislava, 942 48, Slovakia
bInstitute for Astronomy, University of Hawaii, 2680 Woodlawn Drive, Honolulu, HI 96822
cDipartimento di Matematica, University of Pisa, Piazzale B. Pontecorvo 5, 56127 Pisa,
dInstitute of Astronomy, National Central University, No. 300, Jhongda Rd., Jhongli,
eDepartment of Physics, P.O. BOX 64, 00014 University of Helsinki, Finland
Taoyuan, 32001, Taiwan.
Italy
Submitted June 6, 2018
Manuscript pages: 36
Tables: 3
Figures: 9
1
Proposed Running Head: Searching for the first NEO Family
Editorial correspondence to:
Eva Schunova
Institute for Astronomy
University of Hawaii
2680 Woodlawn Drive
Honolulu, HI 96822
Phone: +1 808 294 6299
Fax: +1 808 988 2790
E-mail: [email protected]
2
Abstract
We report on our search for genetically related asteroids amongst the near-Earth
object (NEO) population -- families of NEOs akin to the well known main belt aster-
oid families. We used the technique proposed by Fu et al. (2005) supplemented with
a detailed analysis of the statistical significance of the detected clusters. Their signifi-
cance was assessed by comparison to identical searches performed on 1,000 'fuzzy-real'
NEO orbit distribution models that we developed for this purpose. The family-free
'fuzzy-real' NEO models maintain both the micro and macro distribution of 5 orbital
elements (ignoring the mean anomaly). Three clusters were identified that contain
four or more NEOs but none of them are statistically significant at ≥ 3σ. The most
statistically significant cluster at the ∼ 2σ level contains 4 objects with H < 20 and
all members have long observational arcs and concomitant good orbital elements. De-
spite the low statistical significance we performed several other tests on the cluster to
determine if it is likely a genetic family. The tests included examining the cluster's
taxonomy, size-frequency distribution, consistency with a family-forming event during
tidal disruption in a close approach to Mars, and whether it is detectable in a proper
element cluster search. None of these tests exclude the possibility that the cluster is
a family but neither do they confirm the hypothesis. We conclude that we have not
identified any NEO families.
Key Words: NEAR-EARTH OBJECTS, ASTEROIDS, DYNAMICS
3
1
Introduction
We report here on our null search for a statistically significant cluster of genetically related
near-Earth objects (NEO). The identification of such a NEO 'family' would enable further
research into the physical characteristics of NEOs as was the case with the identification of
asteroid families in the main belt, Jupiter Trojan and Trans-Neptunian minor planet popula-
tions. The physical characteristics of the ensemble of NEO family members -- taxonomic and
mineralogical types, sizes, rotation periods, shapes, pole orientations, existence of satellites,
etc. -- would lead to a better understanding of their morphology and the mechanisms affect-
ing their dynamical and collisional evolution. If the family members have a small minimum
orbital intersection distance (MOID) with the Earth (or any other planet) their presence on
a sub-set of dangerous orbits will increase the total impact probability over what is currently
understood because the enhancement in orbit element phase space is not incorporated into
contemporary NEO models.
In one sense there are already some known NEO families since associations have been
proposed between several meteor showers and their presumed parent body NEOs (both
asteroids and comets, e.g. Sekanina 1973, Porubcan et al. 2004). The most known and
well-accepted is the connection between the unusual B-type NEO (3200) Phaethon and the
Geminid meteor complex Ohtsuka et al. (2008). However, the associations between meteors
and NEOs are not families in the traditional undestanding. The known asteroid families are
produced through the catastrophic disruption of a parent asteroid in a severe impact with
another asteroid. The discovery of genetically related pairs of asteroids in the main belt
(e.g. Vokrouhlick´y & Nesvorn´y 2008) suggests that other mechanisms may produce asteroid
families such as the spin-up and rotational fission of rapidly rotating objects or the splitting
of unstable binary asteroids. Meteoroids are the small-end tail of those family creation
mechanisms but can also be produced through surface ejection driven by the volatilization of
sub-surface ices on comets or 'active asteroids'. While no formal distinction exists between
the meteor 'families' and the asteroid families the size ratio between the largest and second
largest fragments nicely divides the samples and creation mechanisms.
No NEO families have been identified and confirmed. Drummond (2000) suggested that
there were many 'associations' in the NEO population but Fu et al. (2005) showed that all of
them were likely chance alignments of their orbits. The dearth of NEO families is in contrast
to the more than 50 families known in the main belt (e.g. Nesvorn´y et al. 2005). Indeed,
Hirayama (1918) proposed the existence of main belt asteroid families when there were only
790 known main belt asteroids while there are now more than 7,500 known NEOs.
Asteroid families are typically identified by the similarity of the family member's or-
bital elements. Several metrics have been developed to quantify the orbital similarity (e.g.
Southworth & Hawkins 1963, Valsecchi et al. 1999, Jopek et al. 2008, Vokrouhlick´y & Nesvorn´y
2008). The best known is the Southworth-Hawkins D-criterion (DSH; Southworth & Hawkins
1963) that was developed and used to search for parent comets of meteor streams and to
link meteor streams with NEOs (e.g. Gajdos & Porubcan 2005, Ohtsuka et al. 2008).
It is important to keep in mind that simply identifying members within a population with
similar orbital elements does not mean that they are actually a genetically related family.
4
i.e. that they do not necessarily derive from a single parent object. As the number of objects
in the population increases there is a corresponding increase in the likelihood that chance
associations will mimic families and extra care must be taken to establish a proposed family's
statistical significance.
The major difference between the ability to identify families in the NEO and main belt
populations is that the orbits of asteroids in the main belt are stable on timescales comparable
to the age of solar system. Their long term stability allows the calculation of most objects'
proper elements (e.g. Knezevi´c et al. 2002), essentially time-averaged orbital elements, that
are better suited to the identification of similar orbits that are the hallmarks of asteroid
families. Furthermore, the non-gravitational evolution of the main belt asteroids' orbits is
slow under the influence of the Yarkovsky effect (e.g. Vokrouhlick´y et al. 2006a) so that they
occupy the same orbital element phase space for long periods of time.
The main belt families identified by the similarity in their orbit elements range in age
from millions to billions of years. Spectroscopic campaigns have confirmed that the members
of older families typically share the same spectral type; as expected if they all derive from
the same parent body and/or are covered by similar regolith during the family formation
event (e.g. Cellino et al. 2002, Ivezi`c et al. 2002a). Near Earth objects on the other hand
reside in a turbulent dynamical environment with average lifetimes of ∼ 106 yr under the
strong gravitational influence of the terrestrial planets and Jupiter (e.g. Morbidelli et al.
2002, Gladman et al. 1997). Thus, the calculation of NEO proper elements is difficult because
they may cross the orbits of the planets during their time evolution (Gronchi & Milani 2001).
The resulting NEO proper elements are only valid for the time between their close encounters
with the planets and when mean motion resonances of low order with the planets do not
occur.
The search for similar orbits amongst the NEO population using only the osculating
elements is even more limited in the time frame over which the orbits maintain their coherence
(Pauls & Gladman 2005). Furthermore, the non-gravitational accelerations acting on NEOs
are generally much stronger than those acting on the more distant asteroids of the solar
system. Thus, if a NEO family exists it will only be possible to identify it for a short time
period after its creation as its members rapidly disperse.
NEO families might form by several imaginable methods (Fu et al. 2005):
• spontaneous disruption (e.g. YORP spin-up, internal thermal stresses)
• tidal disruption during close approach to a planet
• intra-NEO collisions
• collision with a smaller main belt asteroid
It is difficult or impossible to assign a likelihood to the different formation mechanisms.
While 11% of comets spontaneously disrupt (e.g. Massi & Foglia 2009) as they enter the
inner solar system (1.0AU < q ≤ 1.5AU), presumably due to thermal stresses induced by
approaching the Sun, the NEOs have typically resided in the terrestrial planet zone for long
5
periods of time and many orbital periods. They are thus unlikely to disrupt by this mecha-
nism. On the other hand, simulations suggest that the YORP thermal torque can increase an
asteroid's spin rate to the point where it spontaneously sheds material and some of this ma-
terial may be ejected faster than the parent body's escape speed (e.g. Walsh & Richardson
2006, Pravec et al. 2010, Jacobson & Scheeres 2011).
If the process is repeated multiple
times over a short time span it might be possible to create an asteroid family in this manner.
Families created in this way might be distinguished by their rotation rates, mass ratios, or
pole orientations of their members.
Tidal disruption of NEOs may occur when they pass close to a planet (e.g. Richardson et al.
1998, Walsh & Richardson 2006, 2008) as occurred in the production of the family of objects
associated with Comet Shoemaker-Levy-9 (Sekanina et al. 1994). While the tidal disruption
of asteroids has been simulated under a variety of conditions (e.g., spin rate, pole orienta-
tion, closest approach distance, encounter speed), there is no estimate of how many tidal
disruptions actually occur.
The likelihood of forming NEO families with the third mechanism is probably relatively
small, since the space number-density of NEOs is very small compared to the space number-
density of main belt asteroids (Bottke et al. 1996).
Perhaps the most likely formation mechanism for a NEO family is a collision between the
NEO and a smaller main belt asteroid (Bottke et al. 1996). Many NEOs remain on orbits
with aphelia in or beyond the main belt where they are typically slower than other asteroids
at the same distance and therefore suffer higher speed collisions compared to those between
two main belt asteroids at the same heliocentric distance. Thus, smaller projectiles could be
effective impactors at catastrophically disrupting a NEO.
In summary, NEO families will provide to the opportunity of exploring the physics of
asteroid disruptions at different scales and for different reasons than those observed in the
main belt.
2 Method
We searched for families in the NEO population using the method proposed by Fu et al.
(2005). The technique identifies subsets of objects with similar orbits within a population.
We used their osculating elements, because of the problems involved in calculating NEO
proper elements. We are thus limited to identifying NEO families that have formed relatively
recently though this is not really a problem because 1) NEO dynamical lifetimes1 are on the
order of a million years and 2) young families are interesting.
We used the Southworth-Hawkins DSH criteria (§2.1) to quantify the similarity between
two orbits. It could be argued that other criteria would yield better results but this criteria
has been well tested and forms the basis of the Fu et al. (2005) technique. A set of NEOs
that all have mutual DSH criteria below a threshold Dcluster is considered a 'cluster'. The
technique allows for and favors sub-clustering within the cluster under the assumption that
1A NEO's dynamical lifetime is the time during which it remains in the NEO region.
6
there could be a tight 'core' within a cluster surrounded by a looser assemblage of related
objects. We used the Minor Planet Center (MPC) mpcorb.dat orbit element data set that
contained 7,563 NEOs as of March 2011 (through 2011 DW ). The technique is described in
detail below.
The real difficulty in identifying NEO families is not the identification of the clusters
but in establishing their statistical significance.
Indeed, as shown by Fu et al. (2005) for
NEOs and Pauls & Gladman (2005) for fireballs, it is surprisingly difficult to prove that
similar orbits are statistically significant within a population. We first tested our method
on synthetic family-free NEO orbit models but then developed more realistic family-free
NEO models derived from the known NEO population. We used multiple instances of the
family-free NEO models to establish the statistical significance of our NEO clusters.
2.1 The Southworth-Hawkins D-criterion
The Southworth-Hawkins DSH metric quantifies the similarity of two orbits using five orbital
elements (Southworth & Hawkins 1963). We use the DSH metric in its original form because
it has an established pedigree (e.g. searching for parent comets of meteor streams) that has
been used to successfully identify members of known meteor streams (Southworth & Hawkins
1963, Sekanina 1970) and main belt asteroid families (Lindblad & Southworth 1971).
The DSH metric between two orbits denoted with subscripts m and n is defined by:
with
DSH = pd1
2 + d2
2 + d3
2 + d4
2,
d1 =
qm − qn
AU
,
d2 = em − en,
d3 = 2 sin (I/2),
d4 = (em + en) sin (Π/2)
I = arccos [cos im cos in + sin im sin in cos (Ωm − Ωn)],
Π = (ωm − ωn) ± 2 arcsin(cid:20)cos
im + in
2
sin
Ωm − Ωn
2
sec
I
2(cid:21),
(1)
(2)
(3)
(4)
(5)
(6)
(7)
where q = a(1 − e) is the perihelion distance, a is the semi-major axis, e is the eccentricity, i
the inclination, Ω is the longitude of the ascending node and ω is the argument of perihelion.
I represents the angle between the poles of the two orbits and Π represents the angle between
their perihelia directions (Drummond 2000). Use the positive sign for the arcsin term in eq. 7
when Ωm − Ωn ≤ 180◦ and the negative sign otherwise.
7
2.2
Identifying clusters in orbital element space
We adopted the cluster identification method developed by Fu et al. (2005) that in turn
incorporated the techniques described by Drummond (2000). A 'cluster' is a grouping of
objects with mutually similar orbits whereas we will reserve the term 'family' for a cluster
with high statistical significance that contains members that are likely genetically related.
Fu et al. (2005) showed in limited testing that the method is capable of identifying synthetic
NEO families with minimal contamination.
We grouped objects into clusters based on the values of 4 parameters that are described
in greater detail below:
• Dcluster: The maximum DSH between any pair of objects in a cluster.
• Dpair: The maximum DSH between 'tight' pairs in a cluster.
• SCRmin: The minimum String length to Cluster size Ratio.
• P Fmin: The minimum fraction of pairs of asteroids in the cluster with DSH < Dpair.
We identified candidate clusters as sets of N objects with mutual DSH < Dcluster. Within
each candidate cluster we then identified all pairs (n) of asteroids with DSH < Dpair. The
pair fraction P F is the number of detected pairs divided by the number of all possible pairs
in the cluster (P F = n/ N (N −1)
) and we required that all our clusters satisfy P F > P Fmin.
Then, within each candidate cluster we determined the maximum string length L -- the
maximum number of objects that are connected in a continuous pair-wise fashion such that
each sequential pair in the string satisfies DSH < Dpair. The SCR is the ratio of the number
of objects in the string to the number of objects in the cluster, SCR = L/N, and our final
set of clusters all satisfy SCR > SCRmin. Note that a string must contain more than two
objects.
2
The goal of our selection criteria is to identify tight clusters of objects in orbital element
space but the P F and SCR cuts recognize that as NEOs undergo rapid dynamical and non-
gravitational evolution some of the members may evolve quickly onto different orbits. The
pair and string searches allow for a tight 'core' of objects with a periphery of other objects.
2.3 Selecting thresholds for the cluster identification algorithm
While it is simple to verify that our algorithm can identify objects in clusters with simi-
lar orbital elements it is not trivial to select the four threshold parameters (Dcluster, Dpair,
SCRmin, P Fmin) to maximize the cluster detection efficiency while minimizing the contami-
nation by false positives. This required a family-free NEO orbit distribution model that also
incorporated the observational selection effects typical of the asteroid surveys contributing
the current NEO inventory. The selection effects are important because e.g. they favor the
discovery of objects on Earth-like orbits and therefore increase the orbital element phase-
space density of known objects on these types of orbits.
8
2.3.1 Synthetic family-free NEO model
To generate our synthetic family-free NEO model we started with the set of NEOs from the
Synthetic Solar System Model (S3M; Grav et al. 2011). The S3M includes over 11 million
objects ranging from those that orbit the Sun entirely interior to the Earth's orbit to the most
distant reaches of the solar system. The 268,896 NEOs with absolute magnitude H < 25.0
in the S3M were generated in accordance with the (a, e, i) orbital element residence-time
distribution of the Bottke et al. (2002) NEO model. The model does not include any NEO
families.
To model the observational selection effects on the S3M NEO population we performed
a long-term survey simulation using the Pan-STARRS Moving Object Processing System
(Denneau et al. 2007). MOPS was developed to process source detection data from Pan-
STARRS (Kaiser et al. 2010) but also incorporates real-time processing of synthetic detec-
tions to monitor the system's performance. Thus, it can be also used as a pure survey
simulator.
In an ecliptic longitude (λ0) and latitude (β) system centered on the opposition point
(λ0, β) = (0◦, 0◦) the MOPS simulated survey2 region is broken into two regions covering
about 5,500 deg2: 1) the opposition region with λ0 < 30◦ and β < 40◦ and 2) two 'sweet
spots' with β < 10◦ and 60◦ < λ0 < 90◦.
MOPS uses a full N-body ephemeris determination to calculate the exact (RA, Dec)
of every NEO in each synthetic field and then degrades the astrometry to the realized PS1
astrometric error level of 0.1′′ (Milani et al. (2012) have shown that PS1 achieves ∼ 0.13′′
absolute astrometric error). The photometry for each object is degraded in a S/N-dependent
manner such that as S/N→5 the magnitude error approaches ∼0.1 mag. MOPS then makes
a cut at S/N=5 to simulate the statistical loss of detections near the PS1 system's limiting
magnitude1 of R ∼22.7. Each field is observed twice each night within ∼15 minutes to
allow the formation of tracklets, pairs of detections at nearly the same spatial location, that
might represent the same solar system object. Fields are re-observed 3 times per lunation
(simulated weather permitting) and tracklets are linked across nights to form tracks that are
then tested for consistency using an initial orbit determination (IOD).
Detections in tracks with small astrometric residuals in the IOD are subsequently dif-
ferentially corrected to obtain a final orbit. Our four-year MOPS survey simulation on the
S3M NEO population yielded 8,020 derived objects (synthetic discoveries). Fig. 1 shows
that the derived synthetic NEOs are a rough match to the known NEOs. Thus, our NEO
survey simulation yields a set of synthetic NEOs that are a proxy for the known population
including their observational selection effects (Jedicke et al. 2002). Pan-STARRS is currently
operating a single prototype telescope on Haleakala, Hawaii, known as PS1.
2The simulated survey discussed herein is only loosely related to the actual PS1 survey. The details of
the survey simulation are not important -- all that matters is that the simulated survey reproduces the
observational selection effects of the ensemble of surveys that produced the known NEO population. The
litmus test is whether the resulting simulated orbit element distributions match the known orbit distributions.
9
2.3.2 Cluster identification in the synthetic NEO model
With a large amount of computing time we could identify the set of clusters detectable in
the real NEO population for all possible combinations of the four thresholds -- Dcluster,
Dpair, SCRmin and P Rmin. But this is impractical because we will also need to run orders
of magnitude more tests to establish the detected clusters' statistical significance. We were
guided in our selection of the thresholds by previous work, our own experience, and testing
the algorithm on the synthetic NEO model.
Fig. 2 shows the results of the cluster identification method applied to the family-free
synthetic NEO model using Dcluster = 0.060 and Dpair = 0.058 without any cuts on SCR and
P F . Even though our DSH thresholds are already considerably tighter than those proposed
by Fu et al. (2005) we still identify many false clusters containing 3 members. We even
identify one false 5-member cluster with SCR = 0.8 and P F = 0.3! However, the number of
false clusters drops quickly with the number of cluster members (42 triplets, seven 4-member
clusters and only one 5-member cluster). As the synthetic model is a good representation of
the real NEO population we expect roughly the same number of false clusters amongst the
real NEOs when we use the same Dcluster - Dpair cuts. Thus, even though we will explore
the full range of tighter thresholds on the Dcluster and Dpair values as described below, based
on Fig. 2 we will use SCRmin = 0.75 and P Fmin = 0.5 to identify clusters containing ≥ 4
members and not inspect the clusters containing NEO pairs and triplets. Establishing the
smaller clusters' statistical significance will be difficult or impossible using the techniques
developed here.
In the following section §2.4 we will use the thresholds derived here to identify real NEO
families for detailed analysis. However, when we measure the statistical significance of those
NEO families in §3 we will abandon the synthetic model in favor of a more realistic one --
but the more realistic model must be created using the thresholds derived here.
2.4 NEO clusters in the real population
We identified three clusters of four or more members in the real NEO population from the
mpcorb.dat database3.
(We will ignore the 13 triplets and 243 pairs identified with the
same cuts.) The members' orbital elements and other physical parameters are provided in
Tables 1 and 2. The three clusters are labelled C1, C2 and C3, and have 4, 6 and 5 members
respectively. The absolute magnitudes of the cluster members spans 18.5 < H < 29.5
corresponding to diameters ranging from several meters up to several hundreds of meters
depending on the choice of albedo.
The total number of ≥ 4 member clusters identified in the real data should be compared
to the 0 ≥ 4 member clusters identified in the family-free synthetic data with SCRmin = 0.75
and P Fmin = 0.5 (see Fig. 2). The disparity in the number of ≥ 4-member clusters could be
due to the presence of real NEO families but we will argue below that it is due to the lack
of fidelity in the synthetic NEO model -- we find that the synthetic NEO model must be
3http://www.minorplanetcenter.net/iau/MPCORB.html
10
exquisitely matched to the real NEO population in order to assess the statistical significance
of the detected clusters.
The C1 cluster is the only one composed of objects with H < 20 and, more importantly,
the only one containing objects with orbital arcs longer than 100 days. The C1 cluster
members' orbital element uncertainties are typically 1-2 orders of magnitude lower than the
other two clusters. All the C1 members belong to the Amor4 NEO sub-population so that
they do not cross the Earth's orbit. Indeed, the members of this cluster have a perihelion
distance of ∼ 1.25 AU implying that if the cluster is a genetic NEO family it has probably
never approached close to the Earth or Venus. Considering the similarity of all 5 orbital
elements the cluster must have formed relatively recently and it is unlikely that the cluster
or its members could have approached the Earth and then evolved onto orbits that do not
cross the Earth's on a short time scale.
In contrast to the C1 cluster, the C2 and C3 clusters are composed of small objects with
H > 21.1 and H > 27.7 respectively and include objects with orbital arcs sometimes spanning
just several days. The short arc lengths yield large uncertainties on the orbital elements which
in turn induce a large uncertainty in the C2 and C3 DSH. The clusters thus illustrate how
false associations can arise because of the orbit element uncertainties. The nominal orbits
of the two clusters place them in the Apollo NEO sub-population with the C2 cluster lying
close to the Amor-Apollo transition and the C3 cluster close to the Apollo-Aten transition.
Their location near the transition regions is not a coincidence -- these small objects were
identified by NEO surveys only because their orbits bring them very close to Earth. The
perihelion distance is ∼ 1.00 AU for the members of the C2 cluster while the members of the
C3 cluster are on very Earth-like orbits with a ∼ 1.00 and e < 0.1. Establishing the C2 and
C3 clusters' statistical significance would be difficult because observational selection effects
are not well-characterized for objects in their size range and this induces a large uncertainty
in the orbit and size-distribution models for small NEOs.
Given the problems with the C2 and C3 clusters we will concentrate on establishing the
statistical significance of the C1 cluster. One method of doing so is to run the cluster finding
algorithm on many instances of the synthetic family-free NEO model described in §2.3.1. If
we identify ≤ 3 false clusters of ≥ 4 members in 1,000 realizations of the synthetic NEO
model we could claim that C1 is statistically significant with ≥ 99.7% or ≥ 3σ confidence.
We did not employ this technique because of our concerns with the use of the ∼ 10 year
old Bottke et al. (2002) NEO model that underestimates the number of Amor-type NEOs
like the members of the C1 cluster. There are currently 474 known Amors with H < 18
(as of March 2011) compared to the Bottke et al. (2002) prediction of 310 ± 38 -- a > 4σ
difference between the real and synthetic NEO populations. If the Bottke et al. (2002) NEO
model underestimates the number of Amor-type NEOs then it would imply that we will
overestimate the C1 cluster's statistical significance. Furthermore, the synthetic NEO model
relies on a survey simulation that was not intended to perfectly model real surveys and yields
∼6% more objects than the real NEO population with small but perhaps significant skewing
4The Amors have perihelion distance q in the range 1.0167 AU< q ≤1.3 AU, Apollos have a > 1.0 AU
and q ≤ 1.0167 AU and the Atens have a < 1.0 AU and aphelion Q > 0.983AU.
11
in the synthetic orbital element distributions as shown in Fig. 1. We need a better synthetic
NEO model as described in the next section.
2.4.1 Fuzzy real NEO models
The main problem with the Bottke et al. (2002) synthetic NEO model is illustrated in Fig. 3
-- there is a huge discrepancy between the normalized DSH distributions for the closest pairs
within the real and synthetic NEO populations. But establishing the statistical significance
of our NEO clusters requires a large number of independent high-fidelity family-free NEO
models that incorporate observational selection effects. Thus, we developed NEO models
using a technique that i) maintains both the micro and macro distribution of 5 orbital
elements (ignoring the mean anomaly) and ii) eliminates any possible real clusters.
Our solution was to 'fuzz' the orbital elements of each real NEO around its position in
5-dimensional orbital element space in a manner that maintained the local orbital element
phase-space density and thereby preserves both the intrinsic NEO orbital element distribu-
tion and the observational selection effects. For each NEO (k) in the real population we
identified its closest neighbor as the object with the smallest DSH ≡ Df uzz. We then gen-
erated a new 'fuzzy' synthetic orbit (n) that has DSH ≤ Df uzz with respect to the original
orbit as described below.
If all the difference between the original and new orbit is due to a single orbital element
(e.g. ∆a = an − ak) then we obtain ∆a, ∆e, ∆i, ∆ω, and ∆Ω from:
Df uzz =
(∆a)
AU
(1 − e)
= ∆e p1 + (a/AU)2
= 2 sin(∆i/2)
= 2e sin(∆ω/2)
= qd3
′2 + d4
′2
d3
d4
′ = 2 sin (I∆Ω/2),
′ = 2e sin(Π∆Ω/2),
where
and
I∆Ω = arccos(cid:2)cos2 i + sin2 i cos ∆Ω(cid:3)
Π∆Ω = 2 arcsin(cid:20)cos i sin
2 (cid:21).
I∆Ω
sec
∆Ω
2
(8)
(9)
(10)
(11)
(12)
(13)
(14)
(15)
(16)
(17)
We then generated a 'fuzzed' orbit with (a′, e′, i′, ω ′, Ω′) where the elements x′ were generated
randomly within the range [x − ∆x, x + ∆x]. Finally, we calculated the D∆ between the
12
original and the fuzzed orbit and repeated the generation of the fuzzed orbit for the object
until D∆ < Df uzz. We also repeated the synthetic object generation if the new orbit was not
a NEO (had perihelion a′(1 − e′) > 1.3 AU), was hyperbolic (i.e. e′ ≥ 1), or unphysical (e.g.
a′ ≤ 0, e′ ≤ 0). We generated 1,000 instances of these 'fuzzy-real NEO models' to be used
for establishing the statistical significance of our NEO clusters.
We tested the generation of the fuzzy-real NEO models by generating a series of models
fuzzed by D′
f uzz = ff uzz Df uzz with ff uzz = 0.0, 0.2, 0.5, 1.0, 2.0, and 4.0 and verified that
the models behave as expected and as ff uzz → 0 the generated model reproduces the input
model exactly.
The remaining problem with the fuzzy-real NEO models is that if the real NEO population
contains real families then so will the fuzzy-real NEO models. We needed to remove any
real NEO families from the model first -- but this is difficult to accomplish when there are
no known real NEO families. Instead, we used our own cluster results agnostically with the
assumption that it does not matter whether the clusters we identified are real or not, all that
matters is how often the fuzzing process generates false clusters. Thus, i) using the cluster
detection thresholds determined with the synthetic NEO population described in §2.3.1 we
identified all clusters containing ≥ 3 members and ii) treated the largest member of each
cluster as any other NEO as described above but iii) fuzzed the orbits of the 18 smaller
members of the clusters with ff uzz = 10. We needed to keep the 18 small objects in the
model because they represent about 0.3% of the total known NEO population -- about equal
to the 3σ contribution to the NEO model that we were attempting to measure. On the other
hand we needed to keep them in roughly the correct location in the NEO orbit distribution.
Fig. 1 shows that the fuzzy-real NEO models' semi-major axis, eccentricity and inclination
distributions match the known NEO population far better than the synthetic NEO model.
Even more importantly, Figs. 3 and 4 show that the fuzzy-real NEO models preserve both
the micro and macro DSH distributions (respectively) of the real NEO population that are
critical to using the models to establish the statistical significance of the NEO clusters.
Note that the DSH distribution of the real NEOs is systematically slightly higher than
the fuzzy-real NEO model at small DSH in Fig. 3. This would be the expected signature
if there were real families in the real NEO population. The data point in the lowest DSH
bins in Fig. 3 represent two separate pairs of NEOs. If these two pairs were real it would
imply that they were statistically significant but it turned out that they were subsequently
identified by the Minor Planet Center as corresponding to the same physical object. Thus,
it is reassuring 1) that our fuzzy-real NEO model DSH distribution agrees well with the real
NEOs and 2) that our technique successfully identified identical NEOs.
3 Results and discussion
The shaded regions in Fig. 5 show the results of running our cluster identification algorithm
on the 1,000 family-free fuzzy-real NEO models described above. We searched for clusters
containing ≥ 4 members over the range of Dcluster -- Dpair space with 0.040 ≤ Dcluster ≤ 0.060
where Dpair ≤ Dcluster using fixed SCRmin = 0.75 and P Fmin = 0.5 as described in §2.3.2. It
13
is surprisingly easy to create NEO clusters with ≥ 4 members. The shaded areas in the figure
correspond to regions where we detected a total of a minimum of 3, 46 or 317 clusters in the
1,000 models. The regions correspond to 3-, 2- and 1σ confidence levels on the detection of
clusters in the real NEO population.
Fig. 5 also shows the location of our detected C1, C2 and C3 clusters in the real NEO
population. The clusters can be identified over a limited range of Dcluster -- Dpair with the
most statistically significant point typically having the smallest Dcluster and Dpair. It may be
surprising that the clusters are not identified in a broad fan extending upwards and to the
right in the figure. The truncation in the region in which they are identifiable is due to the
SCRmin = 0.75 and P Fmin = 0.5 cuts -- using looser values of Dcluster and Dpair allows the
cluster to 'absorb' other nearby objects but these objects then typically drive the cluster's
SCR and P F values below the SCRmin and P Fmin thresholds.
Fig. 5 shows that the C1 cluster is statistically significant at ∼2.0σ. The C2 and C3
clusters are likely even less significant based on their large Dcluster and Dpair error bars
extending well into the insignificant region of the figure. It is important to remember that
our calculated statistical significance of the NEO clusters hinges on the reliability of our
fuzzy-real NEO models described in §2.4.1.
Given that all 4 C1 members are relatively large NEOs with H < 20 we repeated the entire
search for clusters using only those NEOs with H < 20 and re-calculated their statistical
significance. As expected, the statistical significance of the C1 cluster increases to about 3σ
because of the reduction in the total number of NEOs to ∼ 3100.
While we have taken every precaution in developing our NEO models we understand that
they have their limitations. For instance, the statistical significance of the NEO clusters
increases/decreases if we 'fuzz' the model more/less (ff uzz >/<1). We assume that the
'natural' value is ff uzz = 1 and use this value when quoting the statistical significance of our
NEO clusters. In the remainder of this work we consider only the C1 cluster as a candidate
NEO family.
3.1 Dismembering the C1 cluster
In this sub-section we perform several tests of the C1 cluster to determine if it is consistent
with being a genetically related NEO family.
3.1.1 Provenance and taxonomy
Despite our minor caveats with the Bottke et al. (2002) NEO model described above it can
still be used to estimate the probability that a NEO derives from one of their five NEO source
regions. The C1 cluster falls in the NEO model bin with central a = 2.150 AU, e = 0.425
and i = 7.5◦ for which objects have ∼35% probability of deriving from the ν6 region and a
∼63% chance of having evolved from the so called 'Mars Crossing' region. In other words, a
nearly 100% probability of deriving from a source in the inner region of the main belt.
The inner region of the belt is dominated by S-class asteroids (e.g. Zellner 1979) so it is a
little surprising that SDSS spectrophotometry (from the 4th Moving Object Catalog, MOC4;
14
Ivezi`c et al. 2002b) for the only available C1 cluster member, 2000 HW23, is not S-like5. We
performed a simple linear interpolation and extrapolation of the 2000 HW23 SDSS MOC4 5-
band spectrophotometry and error bars to the SMASSII bands' central wavelengths (Phase
II of the Small Main-Belt Asteroid Spectroscopic Survey; Bus & Binzel 2002). (The only
extrapolation was from the real data point at 892 nm to the data point at 920 nm.)
Even if at first glance the asteroid appears to be most similar to the O-type due to the
dip in reflectivity at long wavelengths (see Fig. 6) a normalized χ2 fit of the 2000 HW23
spectrophotometry to 26 major asteroid classes from A through Xk (from Bus & Binzel
2002) suggests that it is most consistent with the unusual B-type, a subset of the larger
C-complex. The O-type provides a close second best fit. In general, keeping in mind that
the fit used data extrapolated from the SDSS MOC4 catalogue, 2000 HW23 yields good fits
with the different types in the C-complex -- not the S-class that dominates the inner main
belt.
The case for the C1 cluster being a legitimate family will be strengthened if the spectra
of other C1 cluster members can be shown to be similar to 2000 HW23 and belonging to the
C-complex.
3.1.2 Size-frequency distribution
Under the assumption that C1 is a family we can estimate the slope of the family's size-
frequency distribution (SFD) assuming that the SFD is proportional to 10αH. An unrea-
sonable value of the slope would indicate that the cluster might not be a genetic family.
Converting the observed C1 cluster member's H values into the true SFD requires the de-
tection efficiency as a function of absolute magnitude, ǫ(H), for objects with C1-like orbits.
We estimated ǫ(H) using the synthetic survey simulation described in §2.3.1 by dividing the
number of derived objects with C1-like orbits by the number of objects in the input model.
The survey simulation does a good enough job of reproducing the real NEO population for
the purpose of estimating this efficiency and will not dominate the induced error on the
measured SFD. We found a reasonable fit to the efficiency from the synthetic data using
ǫ(H) =
ǫ0
1 + exp[(H − L)/w]
(18)
with nominal values of ǫ0 = 1.0, L=20.0 and w = 0.76. A maximum likelihood fit to the 4
C1 members' H distribution yields α = 0.23+0.10
−0.08 where the error bars are statistical-only.
The statistical errors on the fit are much larger than the systematic errors introduced by the
uncertainty on the efficiency function parameters.
The slope of the C1 cluster's SFD (Fig. 7) is shallower but not inconsistent with the
overall NEO population's SFD of 0.35±0.02 measured by Bottke et al. (2002) for H<22
and it is surprisingly close to the value of 0.26±0.03 for H&18 from Mainzer et al. (2011).
Parker et al. (2008) measured the SFD for many main belt families and found that they can
have complicated SFDs that can not be fit by a single power-law and that they have a wide
5We used the solar colors from Bilir & Karaali (2005), Allen et al. (2005)
15
range of slopes. The slope varies from 0.35 to 0.97 for single-slope families and from 0.10 to
0.62 at the faint end of the H distribution for families with a broken power law SFD (i.e. H
values approaching those of the C1 cluster members). Furthermore, Richardson et al. (1998)
suggests that tidally disrupted asteroid families have shallow SFDs. Thus, we can not use
the SFD slope to exclude the hypothesis that the cluster is a family since the measured value
is consistent with slopes of established families.
3.1.3 Backward orbit integrations & tidal disruption at Mars
If the C1 cluster is the remnant of a recently formed family then we expect that a backward
integration of the 4 member's nominal orbits would show a rapid convergence to a single orbit
and the time frame for the convergence would indicate the time of the family's formation
(e.g. as observed in the Karin family; Nesvorn´y et al. 2006b). On the contrary, Fig. 8 shows
that the nominal orbits are similar but undergo a gradual dispersal moving 100 ky into the
past.
i.e. they appear to have converged to their most similar orbits at the present time.
The semi-major axis, eccentricity and inclinations of the members remain similar but the
longitude of perihelion and ascending nodes gradually diverge. The effect is summarized in
panel Fig. 8F showing the evolution of the cluster's DSH with time. We note that two pairs
of objects remain tight in their mutual DSH for the entire integration.
Given the apparent stability of the cluster members' orbits we integrated all 4 nominal
orbits backwards for 10 million years and found that they are exceptionally similar back
to 1.5 Myr in the past. The mutual orbital stability is unusual for NEOs given that the
numerical integrations showed numerous close encounters of all objects with Mars during
this time. The lack of convergence of the nominal C1 members' orbital elements suggests
that they are not genetically linked. To firmly establish the statistical significance of the C1
cluster's genetic relationship would require the generation of hundreds of clones of each C1
member and similarly generating clones for a large number of C1-like clusters.
Rather than explore that computationally challenging route we explored a few instances
of the opposite question -- is it possible that a family producing event in the recent past
can reproduce the observed C1 cluster's orbit and DSH distribution? We created 500 clones
for each of the four C1 members where each clone is consistent with the available astrometry
(the clones were generated using the covariance sampling technique in the OpenOrb orbit-
computation package; Granvik et al. 2009). We integrated all the clones backwards in time
for 100 ky and discovered that all of the objects suffered several close approaches with Mars
and a large number of clones approached to within Mars' Hill sphere. Since the C1 cluster
is composed of Amor type NEOs we did not expect nor identify any close approaches to the
Earth and Venus.
One 2000 HW23 clone passed inside Mars' Roche limit ∼70,800 years ago and is therefore
a candidate for tidal disruptions (assuming that the parent body is a spherical fluid or rubble
pile with a density ρ=1.95 g-cm−3 like (25143) Itokawa (Abe et al. 2006), Mars's Roche limit
is ∼3.08 Mars radii). This result needs to be tempered with a comparison to the likelihood
that any NEO will approach to within Mars' Roche limit in the same time frame. We found
that 5,070 of 8,133 known NEO's nominal orbits approach Mars to within 1 Hill radii at least
16
once during the last 100,000 years. Assuming that close approaches to within the Roche limit
are as likely for clones of all NEOs (1 in 2000 C1 clones) then we might expect that about
2.5 of the known NEOs have actually approached to within Mars' Roche limit and could be
candidates for tidal disruption. Perhaps the C1 cluster's hypothetical parent body is one of
them and we will represent it as 2000 HW23 -- the clone that made the closest approach to
Mars mentioned above.
To study the properties of a family of objects created by tidal disruption by Mars we
created twelve 2000 HW23 clones at the moment of its closest approach of 0.63 Mars Roche
radii. The twelve secondary clones have ∆ ~vx, or ∆ ~vy, or ∆ ~vz of ±1 m-s−1 and ±5 m-s−1
with respect to 2000 HW23. Simulations suggest that the tidal disruption process creates
fragments with relative speeds of < 1 m-s−1 (Kevin Walsh, personal communication) so
that our integrations will overestimate the spread in the orbital elements of tidally disrupted
families. All 13 objects (2000 HW23 and its 12 secondary clones) were integrated forward from
the time of closest encounter to the present epoch with the same Bulirsch-Stoer integration
routine and same time step as our other integrations.
Panels A-E in Fig. 9 show that the values and ranges of a, e and i for the ±1 m-s−1
and ±5 m-s−1 groups of secondary clones are comparable to the values and ranges of the
real C1 members. The ranges of Ω and ω, i .e. the spread of values within the two sets of
2000 HW23 clones and the 4 C1 cluster members, are also similar. The values of Ω and ω for
the 2000 HW23 clones and the 4 C1 cluster members do not need to agree at the present time
because the 2000 HW23 secondary clones were created from a single 2000 HW23 clone whose
orbit is slightly different from 2000 HW23 due to the former's deep approach to Mars (their
semi-major axes differ by 0.05 AU). The difference between Ω and ω for the 2000 HW23 clones
and the 4 C1 cluster members is thus due to their different precession rates that are sensitive
to the initial values of semi-major axes (Murray & Dermott 1999). Note that the actual
C1 cluster member's values show evidence of slightly more spreading than the 2000 HW23
secondary clones. Indeed, Fig. 9F shows that the distribution of Dcluster for the 2000 HW23
secondary clones is also tighter than the C1 members. Both observations can be explained
as being due to non-gravitational forces acting on the real objects over the course of the last
∼71 ky.
The distributions of orbital elements and Dcluster in Fig. 9 suggests that the C1 cluster is
consistent with an origin in the tidal disruption of an asteroid during a tidal disruption event
with Mars 70,800 years ago. If true, the C1 cluster is a very young family. In a followup
paper to this one we will show that the lifetime of tidally disrupted families created by Mars
are as long as ∼1 Myr so if they exist it may not be surprising that we are beginning to detect
them. Furthermore, if there is a tidally disrupted asteroid family in the NEO population
it seems more likely to be a C-class asteroid and, ceteris paribus, will disrupt more easily
since they have a lower bulk density than the S-class. (Richardson et al. 1998). Thus, it is
interesting that one C1 member with spectrophotometry, 2000 HW23, may be in the C-class
(see §3.1.1).
17
3.1.4 Proper element cluster search
We also performed a cluster search with NEO proper elements (Gronchi & Milani 2001).
Since the C1 cluster was detected as a cluster in osculating element space it seems plausible
that it should also be detectable in proper element space as described in the introduction. A
complication arises because our cluster search algorithm uses a 5-element DSH-criterion that
incorporates the longitude of the ascending nodes and argument of perihelion (the d3 and d4
terms in eq. 1) while only the proper semi-major axis, eccentricity and inclination (ap, ep, ip)
can be calculated for asteroids. Thus, the DSH for clusters in proper element space, Dp,
must be smaller than those found using the 5-element DSH with the osculating elements --
Dcluster = 0.002 compared to Dcluster = 0.040 used for the osculating element cluster search.
The proper element cluster search revealed that the C1 cluster is not outstanding among
NEO clusters. We identified a C1p cluster in the proper element search that has 3 objects in
common with the C1 cluster -- 2000 HW23, 2001 PF14, and 2008 LN16 -- and there are 166
clusters with 5 or more members that are bound tighter than C1p with Dp=0.018. The C1p
cluster contains two additional objects, 2002 RA182 and 2004 TE18, instead of 2006 JU41. The
osculating element DSH for 2004 TE18 and 2002 RA182 calculated with respect to 2000 HW23
are 0.471 and 0.125, much too large to indicate an orbital similarity.
We found that the stability of the NEOs in the C1 and C1p clusters is not unusual.
We computed the MOID with all the planets from Mercury to Neptune during the secular
evolution of 8,650 NEOs (NEODyS, 29/2/2012) according to the dynamical model used in
the computation of the proper elements (Gronchi & Milani 2001). Roughly 5% of all NEOs
(429) do not cross the trajectory of any planet and ∼ 42% of them (3,636) cross only Mars'
trajectory during their secular evolution.
The six asteroids in the C1 and C1p clusters are of the latter type. Table 3 shows that
all the cluster's members remain far from both the Earth and Jupiter, the major NEO
perturbers, during their entire secular cycles. Of the total 3,636 Mars crossing objects we
find that
• ∼50% of them have min(MOIDEarth) ≥ 0.1 AU, and min(MOIDJ upiter) ≥ 1 AU,
• ∼16% of them have min(MOIDEarth) ≥ 0.2 AU, and min(MOIDJ upiter) ≥ 2 AU.
Thus, even if some of them are affected by mean motion resonances it appears that many Mars
crossing NEOs remain far from close encounters with, and are not affected by perturbations
from, the Earth and Jupiter. A close approach with Mars must be very deep to cause a
significant perturbation.
Thus, there are islands of long-term stability in the NEO orbital element phase space that
might 'collect' NEOs in a manner that could mimic the appearance of a NEO family. The
Bottke et al. (2002) NEO model can not accurately model these regions of stability because
of its low resolution in the orbital element phase space, the relatively small number of objects
that were originally used to map out the NEOs' residence time probability distribution, and
because it is restricted to just 3 of the 6 orbital elements.
18
4 Conclusions
We searched for genetic asteroid families in the NEO population using the method proposed
by Fu et al. (2005) based on identifying clusters of objects with similar orbits. We enhanced
the method's utility by developing a technique for assessing the statistical significance of the
identified clusters using 1,000 realistic family-free fuzzy-real NEO orbital element models.
We created our NEO models by cloning members of the known NEO population in a natural
way based on the 'distance' between each member and its nearest neighbor. The technique
identified three clusters of four or more NEOs among the orbits in the mpcorb.dat data base.
None of the clusters are statistically significant at ≥ 3σ and we conclude that there are as
yet no identified families in the NEO population.
The most statistically significant cluster, C1, contains four objects all with H < 20 and
well-determined orbits with the largest member being asteroid 2000 HW23. We performed
several additional tests of the C1 cluster's family veracity including checking the members'
taxonomic identification, their bias corrected size-frequency distribution, the possibility that
they originated in a family-producing tidal disruption at Mars about 71k years in the past,
and whether it is also identifiable as a cluster in proper element space. None of these tests
exclude the possibility that C1 is a genetically related family but at the same time none of
the tests provide sufficient evidence to elevate the cluster to family status.
The search for families amongst the NEO population is clearly not as straightforward
as the same search amongst the main belt population. Special care must be taken when
assessing the statistical significance of a purported NEO family especially with regard to
accounting for observational selection effects in the NEO population. At the very least,
mere similarity of the orbital elements as evidenced by the DSH criterion being less than
an arbitrary and commonly used value like 0.2 is insufficient for deciding upon any genetic
relationship between NEOs and, by extrapolation, between NEOs and meteors.
19
Acknowledgments
This work was supported by NASA NEOO grant NNXO8AR22G. MG was also funded
by grants #136132 and #137853 from the Academy of Finland. We acknowledge CSC-IT
Center for Science Ltd. for the allocation of computational resources. Schunova's work was
also funded by The National Scholarship Programme of Slovak Republic for the Support of
Mobility of Students, PhD Students, University Teachers and Researchers and VEGA grant
No. 1/0636/09 from the Ministry of Education Of Slovak Republic. We also thank several
colleagues who provided insight into NEO family issues including Kevin Walsh, Giovanni
Valsecchi, Alessandro Morbidelli, Patrick Michel, Joseph Masiero, Davide Farnocchia, and
Seth A. Jacobson.
20
References
Abe, S., Mukai, T., Hirata, N., Barnouin-Jha, O. S., Cheng, A. F., Demura, H., Gaskell, R.,
Hashimoto, T., Hiraoka, K., Honda, T., Kubota, T., Matsuoka, M., Mizuno, T., Nakamura,
R., Scheeres, D. & Yoshikawa, M. (2006), 'Mass and Local Topography Measurements of
Itokawa by Hayabusa', Science 312, 1344 -- 1349.
Allen, D., Rodgers, C. T., Canterna, R., Hausel, E., Flores, E. & Smith, J. A. (2005), Fun-
damental Transformation Equations Between the u'g'r'i'z' and UBVRCIC Filter Systems,
in 'Bulletin of the American Astronomical Society', Vol. 37, p. 1210.
Bilir, S. & Karaali, S. andTuncel, S. (2005), 'Absolute magnitudes for late-type dwarf stars
for Sloan photometry', Astronomische Nachrichten 326, 321 -- 331.
Bottke, W. F., Morbidelli, A., Jedicke, R., Petit, J. M., Levison, H. F., Michel, P. & Metcalfe,
T. S. (2002), 'Debiased Orbital and Absolute Magnitude Distribution of the Near-Earth
Objects', Icarus 156(2), 399 -- 433.
Bottke, W. F., Nolan, M. C., Melosh, H. J., Vickery, A. M. & Greenberg, R. (1996), 'Origin
of spacewatch small Earth-approaching asteroids', Icarus 256, 406 -- 427.
Bus, S. J. & Binzel, R. P. (2002), 'Phase II of the Small Main-Belt Asteroid Spectroscopic
SurveyA Feature-Based Taxonomy', Icarus 158, 146 -- 177.
Cellino, A., Bus, S. J., Doressoundiram, A. & Lazzaro, D. (2002), Spectroscopic properties
of asteroid families, in W. Bottke, A. Cellino, P. Paolicchi & R. P. Binzel, eds, 'Asteroids
III', University of Arizona Press, pp. 633 -- 643.
Denneau, L., J., Kubica, J. & Jedicke, R. (2007), The Pan-STARRS Moving Object Pipeline,
in R. A. Shaw, F. Hill, & D. J. Bell, eds, 'Astronomical Data Analysis Software and
Systems XVI ASP Conference Series #376', Astronomical Data Analysis Software and
Systems XVI ASP Conference Series #376, p. #257.
Drummond, J. D. (2000), 'The D-discriminant and Near-Earth Asteroid Streams', Icarus
146(2), 453 -- 475.
Fu, H., Jedicke, R., Durda, D. D., Fevig, R. & Binzel, R. P. (2005), 'Identifying near-Earth
object families', Icarus 178(2), 434 -- 449.
Gajdos, S. & Porubcan, V. (2005), Bolide meteor streams, in 'Dynamics of Populations
of Planetary Systems, Proceedings of IAU Colloquium #197', IAU Colloquium #197,
pp. 393 -- 398.
Gladman, B. J., Migliorini, F., Morbidelli, A., Zappala, V., Michel, P., Cellino, A., Froeschle,
C., Levison, H. F., Bailey, M. & Duncan, M. (1997), 'Dynamical lifetimes of objects
injected into asteroid belt resonances', Science 277, 197 -- 201.
21
Granvik, M., Virtanen, J., Oszkiewicz, D. & Muinonen, K. (2009), 'OpenOrb: Open-source
asteroid orbit computation software including statistical ranging', Meteoritics and Plane-
tary Science 44, 1853 -- 1861.
Grav, T., Jedicke, R., Denneau, L., Chesley, S., Holman, M. J. & Spahr, T. B. (2011),
'The Pan-STARRS Synthetic Solar System Model: A Tool for Testing and Efficiency
Determination of the Moving Object Processing System', Publications of the Astronomical
Society of the Pacific 123, 423 -- 447.
Gronchi, G. F. & Milani, A. (2001), 'Proper Elements for Earth-Crossing Asteroids', Icarus
152, 58 -- 69.
Hirayama, H. (1918), 'Groups of asteroids probably of common origin', Astronomical Journal
31(743), 185 -- 188.
Ivezi`c, v., Juri`c, M., Lupton, R. & Tabachnik, S.AND Quinn, T. (2002b), SDSS MOC4, in
J. Tyson & S. Wolff, eds, 'Survey and Other Telescope Technologies and Discoveries', Vol.
4836, Proceedings of the SPIE.
Ivezi`c, v., Lupton, R. H., Juri`c, M., Tabachnik, S., Quinn, T., Gunn, J. E., Knapp, G. R.,
Rockosi, C. M. & Brinkmann, J. (2002a), 'Color Confirmation of Asteroid Families', The
Astronomical Journal 124, 2943 -- 2948.
Jacobson, S. A. & Scheeres, D. J. (2011), 'Dynamics of rotationally fissioned asteroids: Source
of observed small asteroid system', Icarus 214, 161 -- 178.
Jedicke, R., Larsen, J. & Spahr, T. (2002), Observational Selection Effects in Asteroid Sur-
veys , in W. Bottke, A. Cellino, P. Paolicchi & R. P. Binzel, eds, 'Asteroids III', University
of Arizona Press, pp. 71 -- 87.
Jopek, T. J., Rudawska, R. & Bartczak, P. (2008), ' Meteoroid Stream Searching: The Use
of the Vectorial Elements', Earth, Moon, and Planets 102(1), 73 -- 78.
Kaiser, N., Burgett, W., Chambers, K., Denneau, L., Heasley, J., Jedicke, R., Magnier,
E., Morgan, J., Onaka, P. & Tonry, J. (2010), The Pan-STARRS wide-field optical/NIR
imaging survey, in L. M. Stepp, R. Gilmozzi & H. J. Hall, eds, 'Ground-based and Airborne
Telescopes III', Vol. 7733, Proceedings of the SPIE, pp. 77330E -- 77330E -- 14.
Knezevi´c, Z., Lemaıtre, A. & Milani, A. (2002), The Determination of Asteroid Proper
Elements, in W. Bottke, A. Cellino, P. Paolicchi & R. P. Binzel, eds, 'Asteroids III',
University of Arizona Press, pp. 603 -- 612.
Lindblad, B. A. & Southworth, R. B. (1971), A Study of Asteroid Families and Streams
by Computer Techniques, in T. Gehrels, ed., 'Physical Studies of Minor Planets', IAU
Colloquium #12, p. 337.
22
Mainzer, A., Grav, T., Bauer, J., Masiero, J., McMillan, R. S., Cutri, R. M., Walker, R.,
Wright, E., Eisenhardt, P., Tholen, D. J., Spahr, T., Jedicke, R., Denneau, L., DeBaun,
E., Elsbury, D., Gautier, T., Gomillion, S., Hand, E., Mo, W., Watkins, J., Wilkins, A.,
Bryngelson, G. L., Del Pino Molina, A., Desai, S., G´omez Camus, M., Hidalgo, S. L., Kon-
stantopoulos, I., Larsen, J. A., Maleszewski, C., Malkan, M. A., Mauduit, J.-C., Mullan,
B. L., Olszewski, E. W., Pforr, J., Saro, A., Scotti, J. V. & Wasserman, L. H. (2011), 'NE-
OWISE Observations of Near-Earth Objects: Preliminary Results', Astrophysical Journal
743, 156.
Massi, G. & Foglia, S. (2009), 'Comet disruption: a statistical approach for the nearly
isotropic comets', Journal of the British Astronomical Association 119, 203 -- 204.
Milani, A., Knezevic, Z., Farnocchia, D., Bernardi, F., Jedicke, R., Denneau, L., Wainscoat,
R. J., Burgett, W., Grav, T., Kaiser, N., Magnier, E. & Price, P. A. (2012), 'Identification
of known objects in solar system surveys', eprint arXiv:1201.2587 .
Morbidelli, A., Bottke, J. W. F., Froeschl´e, C. & Michel, P. (2002), Origin and Evolution of
near-Earth Objects, in W. Bottke, A. Cellino, P. Paolicchi & R. P. Binzel, eds, 'Asteroids
III', University of Arizona Press, pp. 409 -- 422.
Murray, C. D. & Dermott, S. F. (1999), Solar system dynamics, in 'Solar system dynamics',
Cambridge University Press.
Nesvorn´y, D., Enke, B. L., Bottke, W. F., Durda, D. D., Asphaug, E. & Richardson, D. C.
(2006b), 'Karin cluster formation by asteroid impact', Icarus 183(2), 296 -- 311.
Nesvorn´y, D., Jedicke, R., Whiteley, R. J. & Izvezi`c, Z. (2005), 'Evidence for asteroid space
weathering from the Sloan Digital Sky Survey ', Icarus 173(1), 132 -- 152.
Ohtsuka, K., Arakida, H., Ito, T., Yoshikawa, M. & Asher, D. J. (2008), Apollo Asteroid
1999 YC: Another Large Member of the PGC?, in '71st Annual Meeting of the Meteoritical
Society', Annual Meeting of the Meteoritical Society #71, p. 5055.
Parker, A., Ivezi´c, Z., Juri´c, M., Lupton, R., Sekora, M. D. & Kowalski, A. (2008), 'The size
distributions of asteroid families in the SDSS Moving Object Catalog 4', Icarus 198, 138 --
155.
Pauls, A. & Gladman, B. (2005), 'Decoherence time scales for "meteoroid streams"', Mete-
oritics and Planetary Science 40, 1241.
Porubcan, V., Williams, I. P. & Kornos, L. (2004), 'Associations Between Asteroids and
Meteoroid Streams', Earth Moon and Planets 95, 697 -- 712.
Pravec, P., Vokrouhlick´y, D., Polishook, D., Scheeres, D. J., Harris, A. W., Gal Gd, A.,
Vaduvescu, O., Pozo, F., Barr, A., Longa, P., Vachier, F., Colas, F., Pray, D. P., Pollock,
J., Reichart, D., Ivarsen, K., Haislip, J., Lacluyze, A., Kusnir´ak, P., Henych, T., Marchis,
23
F., Macomber, B., Jacobson, S. A., Krugly, Y. N., Sergeev, A. V. & Leroy, A. (2010),
'Formation of asteroid pairs by rotational fission', Nature 466, 1085 -- 1088.
Richardson, D. C., Bottke, W. F. & Love, S. G. (1998), 'Tidal Distortion and Disruption of
Earth-Crossing Asteroids', Icarus 134, 47 -- 76.
Sekanina, Z. (1970), 'Statistical Model of Meteor Streams. II. Major Showers', Icarus 13, 475.
Sekanina, Z. (1973), 'Statistical Model of Meteor Streams. III. Stream Search Among 19303
Radio Meteors', Icarus 18, 253.
Sekanina, Z., Chodas, P. & Yeomans, D. K. (1994), 'Tidal disruption and the appearance of
periodic comet Shoemaker-Levy 9', Astronomy and Astrophysics 289, 607 -- 636.
Southworth, R. B. & Hawkins, G. S. (1963), 'Statistics of meteor streams', Smithsonian
Contributions to Astrophysics 7, 261 -- 285.
Valsecchi, G. B., Jopek, T. J. & Froeschl`e, C. (1999), 'Meteoroid stream identification: a new
approach - I. Theory', Monthly Notices of the Royal Astronomical Society 304(4), 743 -- 740.
Vokrouhlick´y, D., Broz, M., Bottke, W. F., Nesvorn´y, D. & Morbidelli, A. (2006a),
'Yarkovsky/YORP chronology of asteroid families', Icarus 182, 118 -- 142.
Vokrouhlick´y, D. & Nesvorn´y, D. (2008), 'Pairs of Asteroids Probably of a Common Origin',
The Astronomical Journal 136, 280 -- 290.
Walsh, K. J. & Richardson, D. C. (2006), 'Binary near-Earth asteroid formation: Rubble
pile model of tidal disruptions', Icarus 180(1), 201 -- 216.
Walsh, K. J. & Richardson, D. C. (2008), 'A steady-state model of NEA binaries formed by
tidal disruption of gravitational aggregates', Icarus 193(2), 553 -- 566.
Zellner, B. (1979), Asteroid taxonomy and the distribution of the compositional types, in
'Asteroids', University of Arizona Press, pp. 783 -- 806.
24
Name
C1
2000 HW23
2006 JU41
2001 PF14
2008 LN16
C2
1999 YD
2003 UW5
2008 UT5
2008 YW32
2007 YM
2005 WM3
C3
2008 EA9
2010 JW34
1991 VG
2009 BD
2006 RH120
a
∆a
AU ×10−7
e
∆e
×10−6
i
∆i
ω
∆Ω
deg ×10−5 deg ×10−4 deg ×10−4
∆ω
Ω
2.154
2.123
2.120
2.141
2.463
2.470
2.279
2.319
2.584
2.674
1.059
0.983
1.027
1.063
1.033
1
5
4
5
4000
2×105
2×105
5×105
2×107
3×105
342
53
1
107
107
0.424
0.429
0.410
0.421
0.593
0.577
0.555
0.559
0.617
0.620
0.080
0.055
0.049
0.052
0.024
1
2
0.5
0.2
70
300
400
104
3×105
400
38
10
0.2
48
21
7.76
7.61
6.78
7.82
1.38
1.87
1.61
0.99
0.99
1.23
0.42
2.26
1.45
1.27
0.60
7
8
2
2
12
77
91
125
2×105
56
23
42
55
63
9
245
236
254
261
62
52
33
2
11
190
336
43
25
317
10
2
3
3
4
0.2
0.2
0.1
0.6
0.8
0.3
30
7
0.1
10
600
47
52
38
36
10
17
38
71
60
240
129
50
74
253
51
1
2
1
2
20
10
10
40
7×104
10
30
10
0.9
0.06
0.4
Table 1: Orbital elements for the members of 3 NEO clusters (C1, C2 & C3) with four or more
members. Members of the clusters are listed in order of decreasing size. None of the clus-
ters are statistically significant at > 3σ. Data is from the JPL Small-Body Database Browser
http://ssd.jpl.nasa.gov/sbdb.cgi.
25
Name
C1
2000 HW23
2006 JU41
2001 PF14
2008 LN16
C2
1999 YD
2003 UW5
2008 UT5
2008 YW32
2007 YM
2005 WM3
C3
2008 EA9
2010 JW34
1991 VG
2009 BD
2006 RH120
H DS
m
mag
DC DSH ∆DSH
×10−6
m
18.5
19.3
19.5
19.9
21.1
24.3
24.5
25.2
26.2
27.7
27.7
28.1
28.5
28.8
29.5
590
410
370
311
180
40
40
30
180
8
8
7
5
5
3
1,500
1,100
1,000
803
460
110
100
70
40
20
22
18
15
13
9
-
0.041
0.032
0.048
-
0.051
0.044
0.049
0.052
0.026
0.039
0.050
-
0.050
0.049
4
6
4
4
80
3
3
10
1600
170
4
2
270
105
105
Table 2: Absolute magnitude (H), diameter (DS and DC ) and DSH for the members of 3 NEO
clusters (C1, C2 & C3) with four or more members. Members of the clusters are listed in order of
decreasing size. None of the clusters are statistically significant at > 3σ. DSH and its uncertainty
were calculated using the orbital elements and associated errors from www.jpl.org as shown in
Table 1. The diameters DS and DC were calculated using albedos of pV = 0.2 and pV = 0.03
corresponding to the mean albedos of S and C class asteroids as reported by Mainzer et al. (2011).
26
name
min(MOIDEarth) min(MOIDJ upiter)
2000 HW23
2001 PF14
2006 JU41
2008 LN16
2002 RA182
2004 TE18
AU
0.26782
0.27176
0.23183
0.26740
0.27768
0.26278
AU
2.16112
2.23225
2.18723
2.18886
2.16264
2.23812
Table 3: Minimum MOID with the Earth and Jupiter for the NEOs in the C1 and C1p clusters
during the course of their secular evolution.
27
Real NEA population
Synthetic model
Average for 1,000 real fuzzy models
0.0
0.5
1.0
1.5
2.0
2.5
3.0
3.5
4.0
4.5
5.0
5.5
6.0
Semi-major Axis [AU]
0.0
0.0
0.1
0.1
0.2
0.2
0.3
0.3
0.4
0.4
0.5
0.5
0.6
0.6
0.7
0.7
0.8
0.8
0.9
0.9
1.0
1.0
Eccentricity
0.20
0.20
0.18
0.18
0.16
0.16
0.14
0.14
0.12
0.12
0.10
0.10
0.08
0.08
0.06
0.06
0.04
0.04
0.02
0.02
0.00
0.00
0.20
0.20
0.18
0.18
0.16
0.16
0.14
0.14
0.12
0.12
0.10
0.10
0.08
0.08
0.06
0.06
0.04
0.04
0.02
0.02
0.00
0.00
0.30
0.30
0.25
0.25
0.20
0.20
0.15
0.15
0.10
0.10
0.05
0.05
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
0.00
0.00
0
0
10
10
20
20
30
30
40
40
Inclination [deg]
50
50
60
60
70
70
80
80
90
90
Figure 1: Semi-major axis, eccentricity and inclination distributions of the real NEO population
(solid line), the synthetic NEO model (dotted line) and our fuzzy-real NEO model (data points).
28
N=24
N=10
N=2
N=1
N=1
N=0
1.1
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
n
o
i
t
c
a
r
F
r
i
a
P
0.0
0.1
0.2
0.3
0.4
0.5
0.6
0.7
0.8
0.9
1.0
1.1
String-Cluster size Ratio
Figure 2: Pair fraction (P F ) vs. String to Cluster size Ratio (SCR) for false clusters detected in
the family-free synthetic NEO model when Dcluster = 0.060 and Dpair = 0.058. The area of the
filled circles is proportional to the number of false clusters at each P F -SCR combination. Black
circles represent clusters with 3 members, red is for clusters with 4 members, and blue represents
clusters with ≥5 members. Empty black, red and blue circles represent the quantized possible
values for 3, 4 and ≥5 member clusters respectively. It is impossible to have the combinations of
P F and SCR on the left side of the figure.
29
Real NEO population
Synthetic NEO model
Average for 1,000 fuzzy-real NEO models (no discards)
Average for 1,000 synthetic fuzzy NEO models
1E-4
1E-4
1E-4
1E-4
1E-5
1E-5
1E-5
1E-5
1E-6
1E-6
1E-6
1E-6
1E-7
1E-7
1E-7
1E-7
1E-8
1E-8
1E-8
1E-8
s
t
n
u
o
C
d
e
z
i
l
a
m
r
o
N
1E-9
1E-9
1E-9
1E-9
0.00
0.00
0.00
0.00
0.01
0.01
0.01
0.01
0.02
0.02
0.02
0.02
0.04
0.04
0.04
0.04
0.05
0.05
0.05
0.05
0.06
0.06
0.06
0.06
0.03
0.03
0.03
0.03
DSH
Figure 3: DSH distribution of close NEO pairs normalized by the total number of possible pairs
in each population. The solid black histogram represents the real NEO population while the black
data points represent the average±RMS of 1,000 'fuzzy' real NEO models (the 'fuzzing' technique
is described in §2.4.1). The gray histogram represents the nominal synthetic NEO population while
the grey data points represent the average±RMS of 1,000 'fuzzy' synthetic NEO models.
30
10000
10000
1000
1000
100
100
10
10
1
1
t
n
u
o
C
0.1
0.1
0.00
0.00
0.25
0.25
Real NEA population
Average for 1,000 fuzzy-real NEO models
1.25
1.25
1.50
1.50
0.75
0.75
0.50
0.50
Nearest Neighbor DSH
1.00
1.00
Figure 4: DSH distribution for the closest neighboring objects in the real NEO population and the
average±RMS for 1,000 fuzzy-real NEO models.
31
0.060
0.058
0.056
0.054
0.052
0.050
0.048
0.046
0.044
0.042
0.040
>3
3
3
45
2
317
1
C1
C2
C3
r
e
D
t
s
u
c
l
0.040
0.042
0.044
0.046
0.048
0.052
0.054
0.056
0.058
0.060
0.050
Dpair
Figure 5: Statistical significance of the C1, C2 and C3 clusters as a function of the Dcluster and
Dpair thresholds with SCRmin = 0.75 and P Fmin = 0.5. The shaded grey regions mark areas of
3-, 2-, and 1-σ significance for clusters with ≥ 4 members using the fuzzy-real NEO models (see
text for details). The star, circle and triangle represent the most statistically significant location
of the C1, C2 and C3 clusters respectively. The small black lines on the C1 data point represent
the derived errors in Dcluster and Dpair due to the orbital element errors. The long vertical and
horizontal lines through the C2 and C3 cluster's data points represent the same errors for those
clusters. The white region under the diagonal line is unphysical with Dcluster < Dpair.
32
3.0
3.0
2.5
2.5
2.0
2.0
1.5
1.5
1.0
1.0
0.5
0.5
0.0
0.0
-0.5
-0.5
-1.0
-1.0
C
O
B
S
V
e
c
n
a
t
c
e
l
f
e
r
e
v
i
t
l
a
e
R
300
300
400
400
500
500
600
600
700
700
Wavelength [nm]
800
800
900
900
1000
1000
Figure 6: Spectrophotometry of the C1 cluster's largest member, 2000 HW23, from the SDSS
MOC4 by Ivezi`c et al. (2002b) (black points) and the corresponding interpolated and extrapolated
data points and error bars at the SMASSII filters' central wavelengths Bus & Binzel (2002) (grey
triangles and connecting lines). The synthetic SMASSII data points and lines are offset by -0.02 for
clarity. The thin black curves and their adjacent grey curves represent the mean±RMS respectively
of 5 asteroid taxonomic types from the SMASSII survey. All distributions have been corrected to
solar colors and normalized to 1.0 at 550 nm. The relative reflectances of the C, O, B, S and V
types are offset by +1.5, +1.0, +0.5, -1.0 and -1.5 respectively.
33
5
5
4
4
3
3
2
2
1
1
0
0
r
e
b
m
u
n
e
v
i
t
l
a
u
m
u
C
18.4
18.4
18.6
18.6
18.8
18.8
19.0
19.0
19.2
19.2
H [mag]
19.4
19.4
19.6
19.6
19.8
19.8
20.0
20.0
Figure 7: Cumulative absolute magnitude distribution of the four C1 cluster members (grey). The
thick black curve represents the product of the efficiency function (see text) with the result of the
maximum likelihood fit to the SFD with α = 0.23+0.10
−0.08.
34
Figure 8: Panels A-E: Backward evolution of the C1 cluster members' orbital elements from the
present (t0) to 100 ky in the past. Panel F: Dcluster evolution for the C1 cluster from t0 to 100 ky
in the past. Dcluster was calculated with respect to 2000 HW23 and therefore it does not appear on
the panel. Dcluster with respect to 2000 HW23 does not differ significantly from the value calculated
with respect to the average orbit of the C1 members.
35
A
12
10
D
t
n
u
o
C
8
6
4
2
0
2.120
2.125
2.130
2.135
2.140
2.145
2.150
2.155
2.160
0
40
Semi-major axis [AU]
B
7
6
5
4
3
2
1
0
6
5
4
3
2
1
t
n
u
o
C
t
n
u
o
C
t
n
u
o
C
80
160
120
280
Argument of perihelion [deg]
200
240
320
360
14
12
10
E
t
n
u
o
C
0
40
80
120
160
200
240
280
320
360
Longitude of ascending node [deg]
F
8
6
4
2
0
6
5
4
3
2
1
0
0
0.400
0.405
0.410
0.415
0.420
0.425
0.430
0.435
0.440
Eccentricity
C
t
n
u
o
C
8
7
6
5
4
3
2
1
0
6.4
6.6
6.8
7.0
7.4
7.2
Inclination [deg]
7.6
7.8
8.0
8.2
8.4
0.00
0.01
0.02
0.04
0.05
0.06
0.03
Dcluster
Figure 9: Panels A-F: Orbital elements and Dcluster distribution for the C1 custer members (dark
grey) and two sets of 6 secondary 2000 HW23 clones (see text for description). The two sets of clones
have ∆ ~vx, or ∆ ~vy, or ∆ ~vz with respect to 2000 HW23 of ±1 m-s−1 (light grey) and ±5 m-s−1 (white)
respectively at the time of closest Mars approach. There are only 3 values for the real C1 cluster
because Dcluster was calculated with respect to 2000 HW23. The Dcluster values for both sets of 6
secondary clones were calculated with respect to 2000 HW23 that was integrated simultaneously.
36
|
1607.08237 | 2 | 1607 | 2016-10-06T15:37:01 | The population of long-period transiting exoplanets | [
"astro-ph.EP",
"astro-ph.IM"
] | The Kepler Mission has discovered thousands of exoplanets and revolutionized our understanding of their population. This large, homogeneous catalog of discoveries has enabled rigorous studies of the occurrence rate of exoplanets and planetary systems as a function of their physical properties. However, transit surveys like Kepler are most sensitive to planets with orbital periods much shorter than the orbital periods of Jupiter and Saturn, the most massive planets in our Solar System. To address this deficiency, we perform a fully automated search for long-period exoplanets with only one or two transits in the archival Kepler light curves. When applied to the $\sim 40,000$ brightest Sun-like target stars, this search produces 16 long-period exoplanet candidates. Of these candidates, 6 are novel discoveries and 5 are in systems with inner short-period transiting planets. Since our method involves no human intervention, we empirically characterize the detection efficiency of our search. Based on these results, we measure the average occurrence rate of exoplanets smaller than Jupiter with orbital periods in the range 2-25 years to be $2.0\pm0.7$ planets per Sun-like star. | astro-ph.EP | astro-ph | Draft version October 7, 2016
Preprint typeset using LATEX style AASTeX6
With modifications by David W. Hogg and Daniel Foreman-Mackey
THE POPULATION OF LONG-PERIOD TRANSITING EXOPLANETS
Daniel Foreman-Mackey1,2, Timothy D. Morton3, David W. Hogg4,5,6,7,
Eric Agol2, and Bernhard Scholkopf8
[email protected]; Sagan Fellow
2Astronomy Department, University of Washington, Seattle, WA, 98195, USA
3Department of Astrophysics, Princeton University, Princeton, NJ, 08544, USA
4Simons Center for Data Analysis, 160 Fifth Avenue, 7th floor, New York, NY 10010, USA
5Center for Cosmology and Particle Physics, New York University, 4 Washington Place, New York,
NY, 10003, USA
6Max-Planck-Institut fur Astronomie, Konigstuhl 17, D-69117 Heidelberg, Germany
7Center for Data Science, New York University, 726 Broadway, 7th Floor, New York, NY, 10003, USA
8Max Planck Institute for Intelligent Systems Spemannstrasse 38, 72076 Tubingen, Germany
ABSTRACT
The Kepler Mission has discovered thousands of exoplanets and revolu-
tionized our understanding of their population. This large, homogeneous
catalog of discoveries has enabled rigorous studies of the occurrence
rate of exoplanets and planetary systems as a function of their physical
properties. However, transit surveys like Kepler are most sensitive to
planets with orbital periods much shorter than the orbital periods of
Jupiter and Saturn, the most massive planets in our Solar System. To
address this deficiency, we perform a fully automated search for long-
period exoplanets with only one or two transits in the archival Kepler
light curves. When applied to the ∼ 40, 000 brightest Sun-like target
stars, this search produces 16 long-period exoplanet candidates. Of
these candidates, 6 are novel discoveries and 5 are in systems with inner
short-period transiting planets. Since our method involves no human
intervention, we empirically characterize the detection efficiency of our
search. Based on these results, we measure the average occurrence rate
of exoplanets smaller than Jupiter with orbital periods in the range
2–25 years to be 2.0 ± 0.7 planets per Sun-like star.
Keywords: methods: data analysis - methods: statistical - catalogs -
planetary systems - stars: statistics
6
1
0
2
t
c
O
6
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
7
3
2
8
0
.
7
0
6
1
:
v
i
X
r
a
2
Foreman-Mackey, Morton, Hogg, et al.
1. INTRODUCTION
Data from the Kepler Mission (Borucki et al. 2011) have been used to discover
thousands of transiting exoplanets. The systematic nature of these discoveries and
careful quantification of survey selection effects, search completeness, and catalog
reliability has enabled many diverse studies of the detailed frequency and distribution
of exoplanets (for example, Howard et al. 2012; Petigura et al. 2013; Foreman-Mackey
et al. 2014; Dressing & Charbonneau 2015; Burke et al. 2015; Mulders et al. 2015).
So far, these results have been limited to relatively short orbital periods because
existing transit search methods impose the requirement of the detection of at least
three transits within the baseline of the data. For Kepler, with a baseline of about four
years, this sets an absolute upper limit of about two years on the range of detectable
periods. In the Solar System, Jupiter – with a period of 12 years – dominates the
planetary dynamics and, since it would only exhibit at most one transit in the Kepler
data, an exo-Jupiter would be missed by most existing transit search procedures.
Before the launch of the Kepler Mission, it was predicted that the nominal mission
would discover at least 10 exoplanets with only one or two observed transits (Yee &
Gaudi 2008), yet subsequent searches for these signals have already been more fruitful
than expected (Wang et al. 2015; Uehara et al. 2016). However, the systematic study
of the population of long-period exoplanets found in the Kepler data to date has been
hampered due to the substantial technical challenge of implementing a search, as well
as the subtleties involved in interpreting the results. For example, false alarms in the
form of uncorrected systematics in the data and background eclipsing binaries can
make single-transit detections ambiguous.
Any single transit events discovered in the Kepler light curves are interesting in their
own right, but the development of a general and systematic method for the discovery
of planets with orbital periods longer than the survey baseline is also crucial for the
future of exoplanet research with the transit method. All future transit surveys have
shorter observational baselines than the Kepler Mission (K2, Howell et al. 2014; TESS,
Ricker et al. 2015; PLATO, Rauer et al. 2014) and given suitable techniques, single
transit events will be plentiful and easily discovered. The methodological framework
presented in the following pages is a candidate for this task.
A study of the population of long-period transiting planets complements other
planet detection and characterization techniques, such as radial velocity (for example
Cumming et al. 2008; Knutson et al. 2014; Bryan et al. 2016), microlensing (for example
Gould et al. 2010; Cassan et al. 2012; Clanton & Gaudi 2014; Shvartzvald et al. 2016),
direct imaging (for example Bowler 2016), and transmission spectroscopy (for example
Dalba et al. 2015). The marriage of the radial velocity and transit techniques is
particularly powerful as exoplanets with both mass and radius measurements can be
used to study planetary compositions and the formation of planetary systems (for
example Weiss & Marcy 2014; Rogers 2015; Wolfgang et al. 2016). Unfortunately
the existing catalog of exoplanets with measured densities is sparsely populated at
The population of long-period transiting exoplanets
3
long orbital periods; this makes discoveries with the transit method at long orbital
period compelling targets for follow-up observations. Furthermore, even at long orbital
periods, the Kepler light curves should be sensitive to planets at the detection limits
of the current state-of-the-art radial velocity surveys.
There are two main technical barriers to a systematic search for single transit
events. The first is that the transit probability for long-period planets is very low;
scaling as ∝ P −5/3 for orbital periods P longer than the baseline of contiguous
observations. Therefore, even if long-period planets are intrinsically common, they
will be underrepresented in a transiting sample. The second challenge is that there
are many signals in the observed light curves caused by stochastic processes – both
instrumental and astrophysical – that can masquerade as transits. Even when the
most sophisticated methods for removing this variability are used, false signals far
outnumber the true transit events in any traditional search.
At the heart of all periodic transit search procedures is a filtering step based on
"box least squares" (BLS; Kov´acs et al. 2002). This step produces a list of candidate
transit times that is then vetted to remove the substantial fraction of false signals
using some combination of automated heuristics and expert curation. In practice, the
fraction of false signals can be substantially reduced by requiring that at least three
self-consistent transits be observed (Petigura et al. 2013; Burke et al. 2014; Rowe et al.
2015; Coughlin et al. 2016).
Relaxing the requirement of three transits requires a higher signal-to-noise threshold
per transit for validating candidate planets that display only one to two transits. Higher
signal-to-noise allows matching the candidate transit to the expected shape of a limb-
darkened light curve, as well as ruling out various false alarms. This is analagous to
microlensing surveys, for which a planet can only be detected once, thus requiring
high signal-to-noise for a reliable detection (Gould et al. 2004).
Recent work has yielded discoveries of long-period transiting planets with only one
or two high signal-to-noise transits identified in archival Kepler and K2 light curves by
visual inspection (Wang et al. 2013; Kipping et al. 2014b; Wang et al. 2015; Osborn
et al. 2016; Kipping et al. 2016; Uehara et al. 2016). These discoveries have already
yielded some tantalizing insight into the population of long-period transiting planets
but, since these previous results rely on human interaction, it is prohibitively expensive
to reliably measure the completeness of these catalogs. As a result, the existing catalogs
of long-period transiting planets cannot be used to rigorously constrain the occurrence
rate of long-period planets.
In this paper, we develop a systematic method for reliably discovering the transits of
large, long-period companions in photometric time series without human intervention.
The method is similar in character to the recently published fully automated method
used to generate the official DR24 exoplanet candidate catalog from Kepler (Mullally
et al. 2016; Coughlin et al. 2016). Since the search methodology is fully automated,
we can robustly measure the search completeness – using injection and recovery
tests – and use these products to place probabilistic constraints on the occurrence
4
Foreman-Mackey, Morton, Hogg, et al.
rate of long-period planets. We apply this method to a subset of the archival data
from the Kepler Mission, present a catalog of exoplanet candidates, and estimate the
occurrence rate of long-period exoplanets. We finish by discussing the potential effects
of false positives, evaluating the prospects for follow-up, and comparing our results to
other studies based on different planet discovery methods.
2. A FULLY AUTOMATED SEARCH METHOD
To find long-period exoplanets in the Kepler light curves, we search for individual,
high signal-to-noise transit signals using a fully automated procedure that can be
broken into three main steps:
1. an initial candidate search using a box-shaped matched filter,
2. light curve-level vetting (using automated model comparison) to remove signals
that don't have a convincing transit shape, and
3. pixel-level vetting to remove some astrophysical false positives.
The following sections describe each of these steps in more detail.
The model comparison step (step 2) is the key component of our method that
enables robust automation but it is also computationally expensive because we must
estimate the marginalized likelihoods of several different models describing a transit
and other processes that "look" like transits but are actually caused by noise. This
step is conservative: unless a signal is a very convincing transit, it won't pass the test.
In practice, this means that all but the highest signal-to-noise events will be rejected
at this step. Therefore, in the inexpensive first step – the initial candidate search – we
can restrict the candidate list to high signal-to-noise events without a substantial loss
in detection efficiency.
2.1. Step 1 – Initial candidate events
It is not computationally feasible to run a full model comparison at every conceivable
transit time in the light curve so we must first find potentially interesting events. For
our purposes, "interesting" means high signal-to-noise and previously unknown.
To generate this list, we use a method much like the standard "box least squares"
(BLS; Kov´acs et al. 2002) procedure with a single (non-periodic) box. After masking
any known transits, we filter the PDC light curves (Stumpe et al. 2012; Smith et al.
2012) using a running windowed median with a half-width of 2 days to remove stellar
variability. We then compute the signal-to-noise of the depth of a 0.6 day-long top
hat on a grid of times spanning the full baseline of observations.
In detail, at each proposal time t0, we hypothesize a box-shaped transit with
.
(1)
duration τ
m(t) =
µ − δ, if t − t0 < τ /2
otherwise
µ,
The population of long-period transiting exoplanets
5
Assuming that the uncertainties on the observed fluxes f (t) are Gaussian with known
2, the likelihood function for the mean flux µ and transit depth δ can be
variance σf
analytically computed to be a two-dimensional Gaussian with mean and covariance
given by linear least-squares. This likelihood function provides a natural scalar
objective: the signal-to-noise of the measured depth computed as a function of time.
In principle this scalar is also a function of duration but we only use a single transit
duration because the following steps in this procedure are only sensitive to transits
with very high signal-to-noise, and in practice, the final results are insensitive to the
specific choice of duration.
To avoid edge effects, we apodize this detection scalar near any large gaps in the
time series using a logistic function with width equal to one transit duration. Finally,
we estimate the background noise level in the detection scalar time series using a
robust running windowed variance estimate of the detection scalar. We accept peaks
that are more than 25-times this background noise level as candidates.
For the Kepler light curves, this procedure yields at least one candidate event
in about 1 percent of targets. For these targets, we investigate the three highest
signal-to-noise events in the following step.
2.2. Step 2 – Light curve vetting
In this step of the method, the goal is to discard any signals that are not sufficiently
"transit-like" in shape. This step is similar to the method independently developed and
recently published by the Kepler team (Mullally et al. 2016). To quantify the quality
of a candidate, we perform a model comparison between a physical transit model and a
set of other parameterized models for systematics. In order for a candidate to pass this
vetting step, the transit model must be "preferred" to any other model as measured
using the Bayesian Information Criterion (BIC). The BIC is not the optimal choice for
this model comparison, but it is more computationally tractable than the alternatives,
such as computing thousands of precise marginalized likelihoods or expected utilities
for each model. The BIC can be efficiently computed and it exhibits the desired
behavior – decreasing with increasing likelihood but flexible models are penalized – and
we find that it performs sufficiently well in practice.
For up to three candidate transit times per light curve, we select a contiguous
chunk of PDC light curve approximately centered on the proposed transit with no
more than 500 cadences (about 10 days) and compute the BIC of each model for this
data set. The BIC for a model k in the set of K models is given by
(2)
where the likelihood function L is evaluated at its maximum, J is the number of free
parameters in the model, and N is the number of data points in the data set.
BICk =−2 lnL∗ + J ln N
For each model, we describe the data using a Gaussian Process (GP; Rasmussen &
Williams 2006) with a Mat´ern-3/2 covariance and mean given by the chosen model
mk(t; θ) parameterized by the parameter vector θ.
6
Foreman-Mackey, Morton, Hogg, et al.
We consider the following mean models (this list provides a qualitative justification
for each model):
• transit – a limb-darkened, exposure-time integrated transit light curve,
• variability – a pure GP model to capture stellar variability,
• outlier – a single outlier to account for a bad data point,
• step – a step function to describe sudden pixel sensitivity dropouts (SPSDs; for
example Christiansen et al. 2013), and
• box – a box to catch signals that are well fit by the search scalar but insufficiently
transit-like to be convincing.
The functional forms of these models are given in Appendix A and the details of the
technical methodology of GP fitting are described in Appendix B.
Figure 1 shows representative events that fall into different classes and the corre-
sponding maximum likelihood model. For each candidate event, the BIC of each of
these models is computed and the event is only passed as a candidate if the transit
model is preferred to all the other models. The box model is the most restrictive com-
parison, vetoing about half of the candidate events in the Kepler light curves, followed
by the variability model. To further restrict to non-grazing transits, we also reject
events where the maximum likelihood impact parameter is greater than 1 − RP/R(cid:63).
Since the search procedure described here was tuned to discover transit signals,
we do not consider the distribution or potential astrophysical nature of any models
besides the transit model. In the future, it would be interesting to relax this goal
and investigate the other model classes; in particular, the box model is sensitive
to astrophysical phenomena, notably occultations of white dwarfs.
In a cursory
investigation it is clear that the majority of signals labeled box in our analysis are
noise; however, a subset are likely to be astrophysical in nature.
The reliability of this method of automated vetting is limited by the specific models
selected in this step. We find that these are sufficient for the targets discussed below
but different target lists or data sets might require additional models to be included
for robust selection.
2.3. Step 3 – Pixel-level vetting
To minimize contamination from background eclipsing binary systems, we require
candidate events to pass a centroid shift test similar to the one used in the official
Kepler transit search pipeline (Bryson et al. 2013). To measure the centroid shift, we
model the flux-weighted centroid traces independently in each coordinate as a multiple
of the best-fit transit model and a GP noise model. By properly normalizing the
transit model, we measure the in-transit centroid shift ∆centroid in pixels. We reject
any candidate event where the estimated transit location is more than half a pixel
The population of long-period transiting exoplanets
7
Figure 1. Representative examples of candidate events flagged by the initial search. Each
example falls into a different model category and the figure shows the data as black points and
the best fit mean model prediction. The examples represent the following model categories:
(a) step (b) variability, (c) box, and (d) transit.
−40−2002040hourssinceevent−1.2−0.60.00.61.2relativeflux[ppt](a)stepKIC8631697−40−2002040hourssinceevent−2024(b)variabilityKIC7220674−40−2002040hourssinceevent−0.50−0.250.000.25relativeflux[ppt](c)boxKIC9411471−40−2002040hourssinceevent−1.8−1.2−0.60.0(d)transitKIC85052158
Foreman-Mackey, Morton, Hogg, et al.
from the out-of-transit centroid
∆centroid
(cid:19)
− 1
(cid:18)1
δ
> 0.5
(3)
where δ is the observed transit depth (Bryson et al. 2013).
3. RESULTS: A CATALOG OF LONG-PERIOD TRANSITING EXOPLANET
CANDIDATES
To limit the scope of this paper while still demonstrating the applicability of our
method, we search the Kepler archival light curves of the brightest and quietest
Sun-like stars for long-period transiting exoplanets. In this section, we describe the
target selection process and the parameter estimation procedure.
3.1. Target selection
We select the ∼ 40, 000 brightest and quietest G and K dwarfs from the Kepler
catalog using the most recent catalog of stellar parameters1 and the cuts used by
Burke et al. (2015):
• 4200 K ≤ Teff ≤ 6100 K,
• R(cid:63) ≤ 1.15 R(cid:12),
• data span ≥ 2 years,
• duty cycle ≥ 0.6,
• Kp ≤ 15 mag, and
• CDPP7.5 hrs ≤ 1000 ppm.
We continue by excluding the light curves of known eclipsing binaries2 (Kirk et al. 2016),
other known false positives (Coughlin et al. 2016), a planet with known transit timing
variations (Kepler-9), and four especially noisy stars (KIC 4482348, KIC 4450472,
KIC 5438845, and KIC 10068041). The final catalog contains 39,036 targets and the
parameter distribution is shown in Figure 2.
Since these data have already been searched for short-period planets, we assume
that all high signal-to-noise candidates with three or more transits have been previously
found (Coughlin et al. 2016). To remove these candidates from consideration, we
mask the cadences within two transit durations of the time when a short-period planet
candidate is known to transit3.
1 Parameters from the q1 q17 dr24 stellar table from the NASA Exoplanet Archive (Huber et al.
2014, with updates).
2 http://keplerebs.villanova.edu/
3 We specifically use the q1 q17 dr24 koi
exoplanetarchive.ipac.caltech.edu/.
from the NASA Exoplanet Archive http://
The population of long-period transiting exoplanets
9
Figure 2. The distribution of stellar parameters for Kepler targets selected for this search
(orange) compared to the distribution of the full Kepler target catalog (black).
3.2. Parameter estimation
For each transit candidate, we constrain the physical parameters of the system by
fitting a section of light curve around each transit using an exposure-time integrated
Keplerian orbit with a quadratic limb darkening law for the central body Foreman-
Mackey & Morton (2016). It has previously been established that the orbital period of
a transiting planet with only one transit can still be constrained given a measurement
of the stellar density and an assumption about the orbital eccentricity (for example
Wang et al. 2015; Osborn et al. 2016). Qualitatively this works because the transit
of a bound body cannot have an arbitrary period for a given duration. This is the
same argument used to justify the "photoeccentric effect" (Dawson & Johnson 2012)
and the method of "asterodensity profiling" (Kipping et al. 2014a). In particular, this
suggests that the periods of single transits in systems with multiple inner planets will
be especially well constrained (Kipping et al. 2012). In this paper, we do not take
advantage of the extra constraints provided by the inner planets, instead treating each
long-period transiting system in isolation, but this would be a good follow-up project.
In the following paragraphs, we describe the components of the probabilistic model
used to infer the planet candidates' properties. To perform parameter estimation
450060007500Teff3.64.04.44.8logg10
Foreman-Mackey, Morton, Hogg, et al.
under this model, we use the Markov Chain Monte Carlo (MCMC) package emcee4
(Foreman-Mackey et al. 2013) with an ensemble of 40 walkers. We run each chain
until at least 750 independent samples – in most cases, we actually produce thousands
of independent samples – are obtained5 and discard the first third of the chain as
burn-in. The posterior constraints on a few physical parameters for the single transit
candidate in the light curve of KIC 8505215 are shown in Figure 3 and all the chains
are made available online6.
Priors - For each candidate in our sample, we take the constraints on the stellar
parameters from the Kepler DR24 stellar properties catalog and assume an empirical
beta function prior on the eccentricities based on the observed eccentricity distribution
of long-period planets discovered using radial velocities (Kipping 2013a). Table 1 lists
all the fit parameters and their prior distributions. Besides these listed priors, we add
the extra constraint that no other transits can occur in the baseline of the Kepler
observations. This constraint is overly conservative because there is some probability
that a second transit could occur in a data gap but we find that, in practice, most of
the posterior mass is at longer periods and the period inferences are not significantly
affected.
Likelihood function - As above, we model the light curve as a Gaussian Process (GP)
with a physical transit model as the mean, and a covariance matrix described by
a Mat´ern-3/2 kernel function. The full likelihood function and some details of GP
regression are given in Appendix B. For computational efficiency, we first perform a
joint optimization of the physical parameters and GP hyperparameters to find the
maximum a posteriori model then keep the hyperparameters fixed and run MCMC
sampling for the 11 physical parameters alone.
4. CATALOG OF TRANSIT CANDIDATES
Applying the search procedure described in Section 2 to the Kepler light curves of
the 39,036 targets selected in Section 3.1, we find 16 convincing transit candidates.
Visual inspection of each candidate confirms the reliability of the classification and
no candidates are manually removed from the catalog. Of these, three candidates
have two transits in the Kepler baseline and the remainder have only one observable
transit. The candidates and their inferred physical parameters are listed in Table 2
and the light curves are plotted in Figure 4. The inferred radius and orbital periods of
the candidates are compared to the short-period Kepler sample and the Solar System
in Figure 5.
4 http://dfm.io/emcee
5 The integrated autocorrelation time is estimated using a robust iterative method as suggested by
Alan Sokal: http://www.stat.unc.edu/faculty/cji/Sokal.pdf.
6 http://dx.doi.org/10.5281/zenodo.58273
The population of long-period transiting exoplanets
11
Figure 3. The posterior constraints on the physical parameters for the single transit
candidate found in the light curve of KIC 8505215. The contour plots show estimates of the
two-dimensional marginalized probability densities and the histograms show the marginalized
density for each parameter. This figure was generated using corner.py (Foreman-Mackey
2016).
Two of the shortest period candidates – both with two observed transits – have
previously been studied in detail (KIC 8800954 and KIC 3239945; Kipping et al.
2014b, 2016). Table 2 indicates the candidates that were also discovered by earlier
searches for long-period transiting systems using visual inspection (Wang et al. 2015;
Uehara et al. 2016). The consistency between our results and the earlier catalogs is
reassuring. In the light curves of targets with previously known short-period planets,
−0.65−0.60−0.55−0.50log10R/RJ0.20.40.6b1.01.52.0log10P/years0.20.40.60.8e−0.65−0.60−0.55−0.50log10R/RJ0.20.40.6b0.20.40.60.8e12
Foreman-Mackey, Morton, Hogg, et al.
Table 1. The inferred parameters and priors used in the inference
name
mean flux
stellar massa
stellar radiusa
limb darkening
symbol
log f(cid:63)
M(cid:63)
R(cid:63)
q1
q2
planet radius
log RP
units
···
M(cid:12)
R(cid:12)
···
···
R(cid:12)
reference time
semi-major axis
& inclination
eccentricity
√
√
√
√
t0
days
a sin i R(cid:12)1/2
a cos i R(cid:12)1/2
···
···
e sin ω
e cos ω
prior
log f(cid:63) ∼ U(−1, 1)
M(cid:63) ∼ N (M(cid:63),cat, σM,(cid:63),cat)
R(cid:63) ∼ N (R(cid:63),cat, σR,(cid:63),cat)
q1 ∼ U(0, 1)
q2 ∼ U(0, 1)
log RP ∼ U(−10, 2)
t0 ∼ U(tcand − 0.5, tcand + 0.5)b
√
√
a sin i ∼ U(−103, 103)/
a
√
√
a cos i ∼ U(0, 103)/
e ∼ β(1.12, 3.09)c
ω ∼ U(−π, π)
a
aStellar parameters and uncertainties taken from the Kepler catalog (Huber
et al. 2014)
b The reference time is constrained to be within half a day of the candidate
transit time
c Kipping (2013b)
Note-There is one further constraint that complicates these priors: the
period of the orbit must be longer than some minimum period Pmin set
by the transit time and the full baseline of Kepler observations.
our automated search did not find any candidates that weren't previously detected
by visual inspection (Uehara et al. 2016) and one candidate (KIC 3230491) reported
by the human analysis was discarded as grazing by our search. The Planet Hunters
citizen science project (Fischer et al. 2012) reported five long-period candidates with
one or two observed transits in our target list (Wang et al. 2015). Of these, we
also find two (KIC 8410697 and KIC 10842718) although we find a second transit in
the KIC 8410697 system that was previously missed. We do not recover the three
other candidates reported by Wang et al. (2015): KIC 5536555, KIC 9662267, and
KIC 12454613. The transits of these candidates are all low signal-to-noise and they
do not pass our initial signal-to-noise threshold. Six of the candidates in Table 2 have
not been previously published.
Of the 16 candidates, 5 have known inner planets with three or more observable
transits (Coughlin et al. 2016). Given the fact that only 844 of the 39,036 targets had
previously known planets, this means that systems with short-period transiting planets
are nearly a factor of 20 more likely to host long-period planets accessible by our
method than systems with no known inner transiting planets. This difference cannot
be accounted for by differences in completeness between targets with known planets
The population of long-period transiting exoplanets
13
and without because the average detection efficiency for both populations is consistent
within sampling uncertainty. Qualitatively, this suggests that these long-period planets
occur with a higher frequency in multi-planet systems or are preferentially aligned
with the plane of any inner planets but a more detailed analysis would be needed to
make a quantitative statement (see, for example, Tremaine & Dong 2012; Fang &
Margot 2012; Ballard & Johnson 2016; Moriarty & Ballard 2015).
The candidate in the light curve of KIC 4754460 is an individual transit candidate
but another deeper eclipse can be found at a Kepler Barycentric Julian Date (KBJD)
of 1587.13; right at the beginning of Quarter 17. This eclipse was missed by the
automated search because only the second half of the eclipse is observed. The most
likely explanation of this system is that the listed candidate is the secondary eclipse
of a binary system but we will keep the candidate in the list and treat this effect
statistically in Section 7.
Five candidate transit events in the light curves of four targets were rejected because
of a significant centroid shift or a large impact parameter. These events are probably
astrophysical eclipses from binary star systems that were not found by previous studies
of long-period eclipsing binary systems. We do not consider these events further in
the following analysis but Table 3 lists these events and their properties for posterity.
5. EMPIRICAL SEARCH COMPLETENESS
To measure the completeness of the search procedure described in Section 2, we
exploit the fact that transit signals are sparse and rare. Therefore, most light curves
contain no transits and we can reliably measure the recovery rate of our method on
synthetic transit signals – with known properties – injected into real light curves.
This procedure is standard practice in the transit literature and it has been used to
determine the completeness of the KOI catalog (Christiansen et al. 2013, 2015) and
other independent transit searches (Petigura et al. 2013; Dressing & Charbonneau
2015; Foreman-Mackey et al. 2015).
To reliably capture the full structure of the search completeness function, the
simulations must sample the (high-dimensional) space of all properties that affect
the probability of detecting a transit: the stellar properties (including variability
amplitudes and time scales), the planet's physical properties and orbital elements,
and any observational effects (noise, spacecraft pointing variations, etc.). For the
modest goals of this paper, we only need a robust constraint on the transit detection
efficiency integrated across the target sample but, even so, many simulations per star
are required.
The procedure for measuring the recovery rate of simulated transits is as follows:
1. First, a star is randomly selected from the target list, and the PDC light curve
and stellar properties for that star are loaded.
14
Foreman-Mackey, Morton, Hogg, et al.
†
r
e
l
p
e
K
/
I
O
K
t
e
n
a
l
p
r
P
0
7
7
/
8
0
1
1
7
6
1
/
0
9
4
e
n
o
n
/
9
9
4
1
2
/
3
9
6
1
2
4
/
4
7
2
1
6
9
.
0
3
7
.
0
5
9
.
0
7
9
.
0
5
9
.
0
0
8
.
0
6
9
.
0
7
9
.
0
5
9
.
0
6
−
0
1
×
7
.
8
e
n
o
n
9
8
9
/
/
0
7
8
1
4
7
1
1
1
9
.
0
5
9
.
0
4
5
1
/
5
3
4
0
9
.
0
4
9
.
0
0
0
.
1
9
−
0
1
×
9
.
3
s
e
t
a
d
i
d
n
a
c
t
e
n
a
l
p
o
x
e
g
n
i
t
i
s
n
a
r
t
d
o
i
r
e
p
-
g
n
o
l
e
h
t
r
o
f
s
r
e
t
e
m
a
r
a
p
d
e
r
r
e
f
n
i
e
h
T
.
2
e
l
b
a
T
*
q
e
T
K
9
1
3
4
−
+
9
2
1
3
3
.
.
4
4
−
+
8
.
2
4
1
4
0
6
6
−
+
1
7
1
5
8
4
3
−
+
0
7
1
3
5
1
1
−
+
6
0
2
8
6
2
4
−
+
5
8
3
9
2
1
−
+
3
0
1
9
3
3
4
−
+
7
3
1
4
2
.
.
7
7
−
+
4
.
9
8
1
2
3
3
3
−
+
6
2
1
3
4
4
3
−
+
9
1
1
2
3
2
1
−
+
4
1
1
7
7
4
3
−
+
3
5
1
6
8
3
2
−
+
9
5
1
3
7
4
4
−
+
8
2
1
9
8
3
2
−
+
6
6
1
t
c
a
p
m
i
n
o
i
t
a
r
u
d
i
s
u
d
a
r
s
r
u
o
h
J
R
7
1
1
2
.
.
0
0
−
+
4
2
.
0
0
1
1
7
1
0
.
.
0
0
−
+
7
0
2
.
0
7
8
3
1
0
0
.
.
0
0
−
+
3
9
8
.
0
7
0
3
2
.
.
0
0
−
+
0
6
.
0
1
2
1
1
.
.
0
0
−
+
5
1
.
0
9
7
5
3
0
0
.
.
0
0
−
+
9
8
8
.
0
9
0
1
2
.
.
0
0
−
+
8
2
.
0
0
0
2
3
.
.
0
0
−
+
8
2
.
0
3
7
1
1
.
.
0
0
−
+
8
1
.
0
2
8
7
6
0
0
0
0
.
.
0
0
−
+
9
9
3
6
.
0
6
1
1
2
.
.
0
0
−
+
3
2
.
0
6
3
3
1
.
.
0
0
−
+
8
6
.
0
1
6
3
2
.
.
0
0
−
+
7
4
.
0
1
2
6
7
.
.
0
0
−
+
5
4
.
1
2
1
7
7
7
0
0
.
.
0
0
−
+
2
0
2
.
6
1
4
5
5
5
.
.
0
0
−
+
2
9
.
5
1
0
7
3
3
.
.
0
0
−
+
5
8
.
0
1
0
2
1
1
.
.
0
0
−
+
7
7
.
9
1
4
6
.
.
1
1
−
+
4
.
9
3
6
8
1
1
.
.
0
0
−
+
6
0
.
0
2
7
2
3
6
.
.
0
0
−
+
4
4
.
7
2
3
4
1
1
.
.
0
0
−
+
6
7
.
5
1
2
2
4
4
0
0
.
.
0
0
−
+
9
9
4
.
8
0
7
2
2
.
.
0
0
−
+
1
8
.
1
1
0
3
3
3
.
.
0
0
−
+
9
4
.
9
8
2
3
4
.
.
0
0
−
+
4
8
.
6
1
8
2
7
7
0
0
0
0
.
.
0
0
−
+
7
2
0
6
.
0
2
8
9
6
0
0
.
.
0
0
−
+
4
0
8
.
2
1
7
9
1
1
.
.
0
0
−
+
6
2
.
0
2
2
2
1
.
.
0
0
−
+
1
5
.
0
8
1
3
5
.
.
0
0
−
+
2
9
.
5
3
4
4
4
5
.
.
0
0
−
+
5
7
.
7
1
3
2
9
9
0
0
.
.
0
0
−
+
4
1
5
.
0
9
9
3
3
0
0
.
.
0
0
−
+
6
7
8
.
0
5
6
1
1
.
.
0
0
−
+
7
6
.
0
3
3
8
9
0
0
.
.
0
0
−
+
2
8
2
.
0
8
7
7
0
0
1
.
.
0
0
−
+
8
9
6
.
0
5
0
2
3
.
.
0
0
−
+
4
0
.
1
7
7
1
1
0
0
.
.
0
0
−
+
7
7
2
.
0
4
5
4
4
0
0
.
.
0
0
−
+
5
5
3
.
0
5
5
2
2
0
0
.
.
0
0
−
+
6
8
3
.
0
6
9
3
4
.
.
0
0
−
+
2
2
.
1
3
1
1
2
.
.
0
0
−
+
3
4
.
0
4
7
2
2
0
0
.
.
0
0
−
+
6
6
2
.
0
7
6
3
4
0
0
.
.
0
0
−
+
3
6
1
.
0
5
6
3
6
.
.
0
0
−
+
0
0
.
2
6
6
1
1
.
.
0
0
−
+
4
7
.
0
1
2
1
1
.
.
0
0
−
+
3
8
.
0
0
t
D
J
B
K
4
6
1
9
1
0
0
0
.
.
0
0
−
+
2
2
7
6
.
6
6
7
d
o
i
r
e
p
s
r
a
e
y
4
5
.
.
3
9
−
+
0
.
7
8
9
6
6
0
0
0
0
0
0
.
.
0
0
−
+
4
1
7
8
2
.
0
2
4
6
6
2
2
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
1
2
7
8
2
3
9
.
2
6
6
4
4
0
0
0
0
.
.
0
0
−
+
9
6
3
8
.
6
2
8
7
7
3
3
0
0
0
0
.
.
0
0
−
+
9
8
5
0
.
9
3
0
1
3
3
1
1
0
0
0
0
.
.
0
0
−
+
1
3
2
1
.
2
4
5
3
3
1
1
0
0
.
.
0
0
−
+
7
7
6
.
4
8
7
8
7
1
1
0
0
0
0
.
.
0
0
−
+
2
9
4
0
.
0
4
1
9
9
4
5
0
0
0
0
.
.
0
0
−
+
8
3
5
8
.
7
9
6
4
4
2
2
0
0
0
0
.
.
0
0
−
+
2
5
6
7
.
2
9
4
8
8
1
1
0
0
0
0
0
0
.
.
0
0
−
+
8
4
6
5
3
.
1
9
1
1
1
3
3
2
0
0
0
0
.
.
0
0
−
+
2
0
1
1
.
4
0
6
9
1
2
3
0
0
0
0
.
.
0
0
−
+
6
7
9
5
.
3
9
3
3
4
6
6
0
0
0
0
.
.
0
0
−
+
2
6
5
3
.
4
5
5
5
5
1
1
0
0
0
0
0
0
.
.
0
0
−
+
2
9
8
0
8
.
0
3
8
7
7
4
4
0
0
0
0
.
.
0
0
−
+
4
4
3
2
.
6
2
2
6
8
1
1
0
0
0
0
.
.
0
0
−
+
4
7
6
2
.
7
5
6
8
0
.
1
.
3
1
−
+
9
.
5
2
2
.
.
1
4
−
+
0
.
4
4
3
5
5
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
7
9
0
8
8
6
8
.
2
6
8
3
8
−
+
4
5
4
5
.
.
3
9
−
+
1
.
9
9
0
.
4
.
5
1
−
+
9
.
9
1
2
9
9
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
7
5
9
9
7
2
9
.
1
1
3
.
.
1
3
−
+
3
.
4
8
6
.
.
1
7
−
+
9
.
4
3
3
.
.
1
4
−
+
9
.
4
1
1
.
.
2
8
−
+
5
.
5
3
5
8
6
.
.
0
2
−
+
6
1
.
3
2
6
.
0
.
6
2
−
+
7
.
2
1
3
7
.
.
1
4
−
+
3
.
4
p
K
6
.
4
1
0
.
4
1
9
.
4
1
6
.
3
1
4
.
3
1
6
.
3
1
0
.
3
1
9
.
3
1
4
.
3
1
0
.
4
1
4
.
4
1
4
.
3
1
9
.
1
1
9
.
4
1
6
.
4
1
5
.
4
1
(cid:63)
R
(cid:12)
R
5
6
0
2
.
.
0
0
−
+
5
7
.
0
3
5
0
0
.
.
0
0
−
+
1
7
.
0
2
8
2
3
.
.
0
0
−
+
3
1
.
1
5
3
1
4
.
.
0
0
−
+
0
1
.
1
6
5
1
3
.
.
0
0
−
+
0
0
.
1
3
6
1
4
.
.
0
0
−
+
8
0
.
1
3
6
0
0
.
.
0
0
−
+
1
7
.
0
6
1
0
2
.
.
0
0
−
+
4
0
.
1
4
6
0
0
.
.
0
0
−
+
6
7
.
0
5
4
1
4
.
.
0
0
−
+
2
9
.
0
1
0
1
8
.
.
0
0
−
+
1
9
.
0
5
4
0
0
.
.
0
0
−
+
3
7
.
0
3
8
1
3
.
.
0
0
−
+
4
9
.
0
8
6
0
3
.
.
0
0
−
+
1
9
.
0
4
8
1
3
.
.
0
0
−
+
4
0
.
1
9
9
0
1
.
.
0
0
−
+
7
9
.
0
ff
e
T
K
9
2
3
7
1
1
−
+
3
1
5
5
6
8
0
8
1
−
+
6
8
7
4
8
2
5
7
1
1
−
+
6
6
7
5
2
5
8
5
1
1
−
+
0
5
0
6
2
7
5
5
1
1
−
+
8
1
9
5
3
3
8
5
1
1
−
+
2
9
9
5
2
8
0
9
1
−
+
7
8
0
5
9
1
2
0
1
1
−
+
0
0
0
6
1
0
0
1
1
1
−
+
6
8
2
5
0
9
6
0
1
2
−
+
2
6
7
5
3
9
4
7
1
1
−
+
5
8
1
5
9
3
1
5
1
1
−
+
0
0
5
4
3
4
3
5
1
1
−
+
9
4
7
5
7
5
5
6
1
1
−
+
8
2
6
5
6
9
5
5
1
1
−
+
4
5
7
5
1
8
0
0
1
1
−
+
8
8
6
5
d
i
c
i
k
b
8
0
9
8
1
2
3
c
5
4
9
9
3
2
3
0
6
4
4
5
7
4
0
4
4
1
5
5
6
a
,
c
7
9
6
0
1
4
8
7
5
9
6
2
4
8
b
5
1
2
5
0
5
8
b
5
3
7
8
3
7
8
c
4
5
9
0
0
8
8
7
0
3
6
0
3
9
b
9
5
1
7
8
1
0
1
b
3
2
7
7
8
2
0
1
9
1
3
1
2
3
0
1
8
6
0
2
0
6
0
1
a
8
1
7
2
4
8
0
1
b
4
2
1
9
0
7
1
1
O
K
e
h
T
†
e
h
T
∗
.
r
e
t
e
m
a
r
a
p
h
c
a
e
r
o
f
s
e
l
p
m
a
s
r
o
i
r
e
t
s
o
p
e
h
t
f
o
s
e
l
i
t
n
e
c
r
e
p
h
t
-
4
8
d
n
a
,
h
t
-
0
5
,
h
t
-
6
1
e
h
t
e
t
a
c
i
d
n
i
s
e
i
t
n
i
a
t
r
e
c
n
u
d
n
a
s
e
u
l
a
v
e
h
T
-
e
t
o
N
.
g
o
l
a
t
a
c
)
6
1
0
2
(
.
l
a
t
e
a
r
a
h
e
U
e
h
t
n
i
d
e
d
u
l
c
n
I
.
g
o
l
a
t
a
c
)
5
1
0
2
(
.
l
a
t
e
g
n
a
W
e
h
t
n
i
d
e
d
u
l
c
n
I
a
b
.
s
t
i
s
n
a
r
t
d
e
v
r
e
s
b
o
o
w
t
s
a
h
i
e
t
a
d
d
n
a
C
c
.
t
e
g
r
a
t
e
h
t
r
o
f
r
e
b
m
u
n
r
e
l
p
e
K
e
h
t
,
e
l
b
a
c
i
l
p
p
a
f
i
,
d
n
a
r
e
b
m
u
n
I
.
o
d
e
b
l
a
o
r
e
z
i
g
n
m
u
s
s
a
d
e
t
u
p
m
o
c
s
i
e
r
u
t
a
r
e
p
m
e
t
m
u
i
r
b
i
l
i
u
q
e
The population of long-period transiting exoplanets
15
Table 3. The signals rejected with a centroid shift or large impact parameter
kic id
time
depth duration
reason
KBJD ppt
hours
3230491
315.3
9.0
6342758
553.9
8463272
641.0
10.3
35.5
8463272
1206.7
35.5
7.4
9.9
4.8
4.8
impact
impact
impact
impact
10668646 1449.2
5.7
12.4
centroid
2. Planetary properties are sampled from the distributions listed in Table 4 with
phase uniformly distributed across the baseline of observations. These properties
are re-sampled until the transit is visible in at least one non-flagged cadence.
3. The transit signal induced by this planet is computed and multiplied into the
PDC light curve.
4. The transit search method described in Section 2 – including de-trending and all
automated vetting – is applied to this light curve with the injected transit signal.
5. This candidate is flagged as recovered if at least one transit within one transit
duration passes all the cuts imposed by the automated vetting.
The fraction of recovered simulations as a function of the relevant parameters gives
an estimate of the probability of detecting an exoplanet transit with a given set of
parameters, conditioned on the fact that it transits the star during a time when the
star was being observed by Kepler. We will call this function Qdet,k(w) where w is the
set of all parameters affecting the transit detectability and k is an index running over
target stars.
Figure 6 shows the fraction of recovered simulations as a function of planet radius
and orbital period based on 819,752 injected signals. This figure shows the transit
detection efficiency falling with decreasing planet radius. This is the expected behavior
because the depth (and signal strength) of a transit scales with the area ratio between
the planet and the star. There is also a slight decrease in the completeness to larger
planet radius. This trend is introduced in steps 2 and 3 of the search procedure where
the tuning parameters were chosen to maximize the yield of convincing small transit
discoveries. The decreasing completeness with orbital period is less intuitive because,
on average, the signal strength should increase as the duration of the transit increases.
In this case, this simplistic treatment misses two important factors. First, in step
1 of the search procedure (Section 2.1) only a single transit duration is used and
second, longer transits are less easily distinguished from stellar variability and they
will, therefore, be discarded in the conservative light curve vetting step (Section 2.2).
16
Foreman-Mackey, Morton, Hogg, et al.
Figure 4. Sections of PDC light curve centered on each candidate (black) with the posterior-
median transit model over-plotted (orange). The y-axis shows the relative apparent flux
of the light curve in parts per thousand (ppt). Candidates with two transits are folded on
the posterior-median period. The plots are ordered by increasing planetary radius from the
top-left to the bottom-right.
−0.50−0.250.00flux[ppt]10321319−1.2−0.60.0flux[ppt]10287723−1.6−0.80.0flux[ppt]8505215−0.80.0flux[ppt]6551440−0.80.0flux[ppt]8738735−3−2−10flux[ppt]8800954−4−20flux[ppt]10187159−4−20flux[ppt]3218908−3.0−1.50.0flux[ppt]4754460−5.0−2.50.0flux[ppt]8410697−4−20flux[ppt]10842718−8−40flux[ppt]11709124−16−80flux[ppt]3239945−4−20flux[ppt]8426957−50−250flux[ppt]9306307−80−400flux[ppt]10602068The population of long-period transiting exoplanets
17
Figure 5. The catalog of long-period transiting exoplanet candidates (green points with error
bars) compared to the Kepler candidates (blue points) and confirmed planets (black points;
Morton et al. 2016) found in our target sample, and the Solar System (orange squares). The
thin black error bars to the left of each candidate indicate the minimum period allowed for
each candidate by the prior assumption that no other transit occurred during the baseline
of Kepler observations of the target. The vertical solid line shows the absolute maximum
period accessible to transit searches that require at least three transits in the Kepler data.
This detection efficiency must then be combined with the geometric transit probabil-
ity function and the window function. For the star k, the geometric transit probability
is given by (Winn 2010)
Qgeom,k(w) =
=
R(cid:63),k + R
(cid:20) 4 π2
ak
G M(cid:63),k
1 − e2
1 + e sin ω
(cid:21)1/3 (cid:20)1 + e sin ω
1 − e2
(cid:21)
(R(cid:63),k + R) P −2/3
(4)
(5)
where R is the planet radius, P is the orbital period, e is the orbital eccentricity, ω is
the argument of periastron, R(cid:63),k is the radius of star k, and M(cid:63),k is the star's mass.
All of these parameters are included in w.
In Equation (4), the term (R(cid:63),k + R) takes grazing transits into account. This
might seem counter intuitive because, as part of the search procedure, we rejected
candidates where the maximum likelihood model had a grazing transit. However,
since the measurement of Qdet,k included a cut on the measured impact parameter, the
110100100010000orbitalperiod[days]110planetradius[R⊕]18
Foreman-Mackey, Morton, Hogg, et al.
Qdet,k term already takes this effect into account. In other words, Qdet,k quantifies the
probability that a transit of a given shape will be detected given that it transits at all
and Qgeom,k – the way it is written in Equation (4) – is the marginalized probability
that the system will transit given its physical parameters.
Approximating the window function using a binomial probability of observing at
least one transit, we find (following Burke & McCullough 2014)
if P ≤ Tk
otherwise
1 − (1 − fduty,k)Tk/P
Qwin,k(w) =
Tk fduty,k/P
where fduty,k is the duty cycle and Tk is the full observation baseline for target k.
Combining these detection effects, the total detection efficiency is given by
Qk(w) = Qdet,k(w) Qwin,k(w) Qgeom,k(w)
.
So that our planet candidate catalog can be easily used for other projects, we
also provide an analytic approximation to the relevant integrated detection efficiency
function
(6)
(7)
(cid:90)
K(cid:88)
Qdet(P, R) =
Qdet,k(w) p(w{P, R}) dw{P, R}
(8)
where p(w{P, R}) is the prior distribution of all the parameters except the period and
radius. We find that a good fit to this integrated completeness is given by the function
k=1
Qdet(P, R)≈
min[max[a(P ) b(R), 0], 1]
1 + exp [−k(P ) (ln R/RJ − x(P ))]
where
a(P ) = a1 ln P/ yr + a2 ,
b(R) = b1 ln R/RJ + b2 ,
k(P ) = k1 ln P/ yr + k2 , and x(P ) = x1 ln P/ yr + x2 .
(9)
(10)
(11)
When fit to the set of 819,752 injected transits, the best fit parameters are given in
Table 5 and the approximation is plotted in Figure 7. Note that we do not use this
approximation in the following analysis but instead compute the relevant integrals
using the injection results directly.
6. THE OCCURRENCE RATE OF LONG-PERIOD EXOPLANETS
Using the catalog of exoplanet discoveries (Section 4) and the measurement of
the search completeness (Section 5), we can now estimate the occurrence rate of
long-period exoplanets. To simplify the analysis, we will make the strong assumption
that none of the candidates are astrophysical false positives (the eclipse or occultation
of a stellar mass companion, around either the target star or a faint background star).
We revisit this assumption and discuss its validity in the following section. As a
further simplification, we also neglect the measurement uncertainties on the planet
The population of long-period transiting exoplanets
19
Table 4. Distributions of physical parameters for transit simulations
name
period
radius ratio
impact parameter
eccentricity
limb darkening
aKipping (2013b)
distribution
log P ∼ U(log 2 yr, log 25 yr)
log RP/R(cid:63) ∼ U(log 0.02, log 0.2)
b ∼ U(0, 1 + RP/R(cid:63))
e ∼ β(1.12, 3.09)a
ω ∼ U(−π, π)
q1 ∼ U(0, 1)
q2 ∼ U(0, 1)
Table 5. The fit parameters for the analytic approximation to the completeness function
parameter
a1
a2
b1
b2
value
−0.13
0.95
−0.20
0.90
parameter
value
k1
k2
x1
x2
0.70
3.06
−0.07
−0.91
parameters (including orbital period). This assumption is justified because we are
only making high-level measurements of the mean occurrence rate in bins larger than
the uncertainties.
Assuming a Poisson likelihood, the occurrence rate density in a volume V – defined
as Pmin ≤ P < Pmax and Rmin ≤ R < Rmax – is (see, for example, the Appendix of
Foreman-Mackey et al. 2014)
d2N
ΓV ≡
(12)
where N is the expected number of planets per G/K dwarf, C(··· ) is the number of
detected planets in the volume, and
d ln P d ln R
C(Pmin, Pmax; Rmin, Rmax)
Z(Pmin, Pmax; Rmin, Rmax)
=
(cid:90)
K(cid:88)
Z(Pmin, Pmax; Rmin, Rmax) =
p(w{P, R}) Qk(w) 1[P, R ∈ V ] dw
(13)
where p(w{P, R}) is the prior distribution of all the parameters except the period and
radius and 1[·] is 1 if the argument is satisfied and 0 otherwise. Using the J injections
k=1
20
Foreman-Mackey, Morton, Hogg, et al.
Figure 6. An empirical estimate of the search completeness as a function of planet radius
and orbital period. In each bin, the completeness is estimated by the fraction of recovered
simulations. The projected histograms show the integrated completeness as independent
functions of period and radius.
351020period[years]0.20.51.02.0RP/RJ0.0480.2110.4990.6690.7270.7100.6350.0460.1940.4680.6160.6570.6300.5690.0430.1930.4600.6050.6230.5910.5200.0380.1740.4330.5290.5290.4920.4270.00.30.60.00.30.6The population of long-period transiting exoplanets
21
Figure 7. An analytic approximation to Figure 6 with the same color scale. The contours
indicate the 0.1, 0.3, 0.5, and 0.7 levels.
sampled uniformly in period and radius and other parameters from p(w{P, R}),
J(cid:88)
j=1
Z(Pmin, Pmax; Rmin, Rmax)≈ K V
J
Qkj (w(j))
(14)
where the sum is over all injections in the volume V .
Using the injection results from Section 5 and the catalog of discoveries from
Section 4, we compute the occurrence rate in the period range 2 to 25 years and in
two radius bins between 0.1 and 1.0 RJ. The calculated occurrence rates are listed in
Table 6. Integrating the two bin model in this range, we find an expected occurrence
rate of
N0.1 RJ−1 RJ, 2 yr−25 yr = 2.00 ± 0.72 planets
(15)
per G/K dwarf with radii in the range 0.1 RJ−1 RJ and periods in the range 2 yr−25 yr.
This result is qualitatively consistent with the Solar System where there is one planet –
Jupiter – in this parameter range and Saturn is just outside the range with an orbital
period of 29 years. In Section 8, we compare with similar occurrence rate estimates
from the literature.
The occurrence rates given here should be interpreted with a few caveats in mind.
First, when we inferred the periods of the planets with only one transit, we assumed
that the period was long enough that no other transit occurred during the Kepler
lifetime. This neglects the small but non-negligible posterior probability – less than one
percent for the typical candidate – that a second transit might have occurred in a data
gap. All of the candidates in our catalog are consistent with having periods this long
but the geometric transit probability decreases quickly with orbital period. For the
351020period[years]0.20.51.02.0RP/RJ0.00.20.40.60.81.022
Foreman-Mackey, Morton, Hogg, et al.
Table 6. The occurrence rate density in two radius bins
Rmin [RJ] Rmax [RJ]
rate densitya
integrated rateb
0.4
0.4
1.0
1.0
0.1
0.1
0.45 ± 0.20 (0.36 ± 0.16)
0.18 ± 0.07 (0.16 ± 0.06)
0.24 ± 0.07 (0.22 ± 0.06)
1.57 ± 0.70 (1.26 ± 0.56)
0.42 ± 0.16 (0.36 ± 0.14)
1.41 ± 0.41 (1.29 ± 0.37)
aThe rate density is given by Equation (12) and the value in parentheses is
computed assuming one candidate is a false positive (Equation 17).
b The integrated rate is computed by integrating the rate density over the bin.
Note that the first two rows do not sum to the last row because each row is
computed assuming that the rate density is uniform across the bin.
Note-These values are computed in the period range 2–25 years.
purposes of this paper, we neglect this effect because its rigorous treatment is subtle,
but comment that this would only ever decrease the occurrence rate estimate. Second,
we assume that each planet candidate transits the star that is characterized by Huber
et al. (2014); we assert that each planet does not transit a fainter companion star or a
background star. If the planet does transit a companion star, then the companion star
must be fainter, and hence denser, causing the period to be underestimated. If the
planet transits a background star, it is more likely to be a giant star due to Malmquist
bias, hence the density of the star and period of the planet would be overestimated.
Either of these scenarios has a small probability, so we expect that our population
estimates will stand, while the parameter estimates for individual candidates should be
taken as provisional until more detailed follow-up is carried out, including high-contrast
imaging, high-resolution spectroscopy, and parallax measurements. Third, we assume
that the Huber et al. (2014) parameters are accurate for each star that is transited
by a planet candidate, and that each transit is unaffected by blending. Malmquist
bias, Eddington bias, and metallicity bias may affect the stellar parameters (Gaidos &
Mann 2012), and so we again caution that the individual parameter estimates should
be taken as provisional until more detailed follow-up is completed.
7. ASTROPHYSICAL FALSE POSITIVES
Various configurations of eclipsing binary stars can mimic the signal of a transiting
planet. However, the occurrence rate calculation presented in the preceding section
assumes no astrophysical false positives among the candidates identified in this work.
In this Section we explore the validity of this assumption.
While an eclipsing binary (EB) typically produces a photometric dip much deeper
than a transiting planet, the depth of the signal may be comparable to that of a planet
if the eclipse is grazing, or if the EB only comprises a small fraction of the total light in
the photometric aperture – a so-called blended eclipsing binary (BEB). Additionally,
if a binary star has an eccentric orbit, it may be oriented so as to present only a
The population of long-period transiting exoplanets
23
secondary occultation and not a primary eclipse, causing a shallow and potentially
flat-bottomed photometric dip without an accompanying tell-tale deep primary signal.
To determine to what extent the catalog of detections presented in this work may
contain such false positives, we simulate populations of detected signals to predict how
many we should expect. To accomplish this, we use the Python package exosyspop
(Morton & Foreman-Mackey 2016), which we developed for this purpose and utilizes
the isochrones, vespa, and batman packages (Morton 2015a,b; Kreidberg 2015) for
simulations of stellar populations and their eclipses.
With exosyspop one can define the parameters of a population model and generate
synthetic catalogs according to the model (and the parameters of a survey) very
efficiently. For example, a population of EBs may be defined by a binary fraction,
power-law distributions in mass ratio and period (within given bounds), and a beta
distribution for eccentricity. This population, initialized with a catalog of target stars
(each of which has a duty cycle and total span of observation), may then be "observed,"
returning a catalog of objects detectable via either primary or secondary eclipse
(according to randomly oriented orbital geometries and accounting for observation
duty cycle and data span). This synthetic catalog includes signal-to-noise estimates
of both the primary and secondary eclipses, the number of detected primary and
secondary eclipses, and the trapezoidal shape parameters of each detection (depth,
duration, and ingress-to-duration ratio, as defined in Morton 2012).
In order to predict how many EBs or BEBs we might expect to detect in this
particular search of Kepler data, we first need to choose reasonable parameters for
the binary star population. To do this, we calibrate the population parameters using
the catalog of detected Kepler eclipsing binaries. We find that a binary fraction of
25% between periods of 20 days and 25 years with a log-flat period distribution and
eccentrities distributed according to β(0.8, 2.0) is able to reproduce well both the
number and period distribution of observed Kepler EBs between 20 and 1000 days.
We thus fix these binary star population parameters for our subsequent EB and BEB
simulations.
To simulate synthetic populations of EB detections, we assign binary stars to the
Kepler target list described in Section 3.1 according to the above EB population
parameters. We consider an EB to be detected if it presents fewer than three eclipses
(either primary or secondary, but not both), if the signal-to-noise ratio is >15, and
the duration of the detected eclipse is <2.5 d. In 100 realizations of these synthetic
observations, we see 7.2 ± 2.5 single- or double-eclipsing EB signals.
To simulate BEBs, we assume an exponentially varying background field star
density across the Kepler field, from 0.005 arcsec−2 at a Galactic latitude b = 20 to
0.05 arcsec−2 at b = 5 (matching up well with the simulations of Morton & Johnson
(2011) at many different Galactic latitudes). Each Kepler target star is then assigned
a number of background stars drawn from a Poisson distribution with mean given
by the expected number of stars to be found within a circle of 4 arcsec radius, given
the appropriate density at its Galactic lattitude. We draw the specific background
24
Foreman-Mackey, Morton, Hogg, et al.
stars from a TRILEGAL (Girardi et al. 2005) field star simulation toward the center
of the Kepler field. Binary companions are then assigned to these background stars
according to the same stellar binary population distribution as the EB population
above, and synthetic detected populations are "observed" according to the same rules
(accounting appropriately for the diluted eclipse depths in the Kepler bandpass). In
100 synthetic observations, we see an average of only 0.41 detected BEBs.
These prediction results suggest that we should indeed expect to see some astro-
physical false positives in our search. However, this does not mean that we should
fear that ∼7 of the planet candidates might be EBs. In particular, we note that these
simulations do not include the full vetting procedure described in Section 4, and it
is likely that the three impact-parameter-rejected candidates are EBs and that the
centroid-rejected candidate is a BEB. Thus, we might expect maybe two or three
additional false positives among our planet candidates.
In order to more precisely quantify which of the candidates might indeed be false
positives, we can inspect the synthetic observation simulations in more detail. In
particular, we can analyze the shape distribution of the different scenarios and compare
them with the observed shapes of the actual Kepler detections in order to quantify
the probability that each of them may be a false positive. These distributions and the
observed shape parameters are plotted in Figure 8.
Following the method of Morton (2012) used to compute false positive probabilities
for the regular Kepler KOI catalogs (Morton et al. 2016), we can calculate the posterior
probability for each of our candidates to belong to each of the three scenarios we
consider (EB, BEB, or planet) as follows:
πiLi(cid:80)
j πjLj
Pri =
,
(16)
where the π factors are the hypothesis priors, L are the hypothesis likelihoods, and the
sum over j is over all the hypotheses. In this case, we determine the relative hypothesis
priors from the synthetic observations, using the mean numbers of "observed" EBs
(7.2) and BEBs (0.4), and choosing the expected number of planets to be 12. We
calculate the hypothesis likelihoods using the depth/duration distributions of synthetic
populations of each scenario and evaluating these distributions at the observed depths
and durations of each candidate signal. To estimate the expected shape distribution
of the planet scenario, we define a custom exosyspop population of planets according
to the two-bin population model described by the median posterior values in Table 6
and generate a population of 1000 detected signals.
We list the probability that each candidate is a planet in Table 2. We find that the
two deepest signals in our candidate catalog (9306307 and 10602068) are very likely
to be EBs, though we note that this result may be dominated by the fact that our
planet population is fixed to have a maximum radius of 1 RJ. Most of the rest of the
candidates have false positive probabilities below 10%. We do note that as discussed
in Section 4, 4754460 (for which we calculate a 5% false positive probability) does
The population of long-period transiting exoplanets
25
show a partial deep eclipse right at the end of the Kepler data that indicates that it
is most likely an EB. Apart from this, the expected number of false positives among
the candidates with R < RJ, according to these calculations, is about one.
In light of this result, we demonstrate the sensitivity of our measured occurrence
rates on contamination by computing a second constraint on ΓV for each volume with
one candidate removed. In this case, Equation (12) would be replaced by
ΓV ≡ C(Pmin, Pmax; Rmin, Rmax) − 1
Z(Pmin, Pmax; Rmin, Rmax)
.
(17)
These updated rates are listed in Table 6. In each case, the results are consistent
within the uncertainties but the difference can be used to get a qualitative sense of
the systematic uncertainty introduced by the false positive population.
We note that in the above procedure we have not corrected our predictions for the
fact that our search has explicitly excluded KOIs that host known Kepler EBs-if
any of these excluded systems show fewer than three eclipses and do not present
both primary and secondary eclipses, then they should also should perhaps count
towards the number of EBs we should have expected to find in this survey. However,
as the Kepler EB catalog does not provide information on whether both primary and
secondary eclipses are detected, we neglect this correction. We note that this is a
conservative decision, in the sense that accounting for the effect of excluded EBs on
our predictions would only further decrease the FPP of the planetary signals, as they
would be even less likely to be caused by EBs.
Figure 8. Predicted eclipse shape distributions for the two false positive scenerios and
exoplanet transits (grayscale heatmap). In this figure, the relative normalization of the maps
are arbitrary but the absolute normalization is discussed in Section 7. The green points
show the shape parameters of the long-period exoplanet candidates from Table 2.
8. COMPARISON WITH THE LITERATURE
The population of long-period planets has been previously studied using radial
velocity, microlensing, and direct imaging surveys. These methods all measure the
0.00.51.01.52.0Duration[days]−3.0−2.5−2.0−1.5−1.0log10(depth)EB0.00.51.01.52.0Duration[days]BEB0.00.51.01.52.0Duration[days]Planet26
Foreman-Mackey, Morton, Hogg, et al.
occurrence rate as a function of planet mass instead of radius. Using the transit
method, however, we do not directly measure the mass of the planet. Therefore,
to compare our results with the literature, we must rely on a mass–radius (M–R)
relationship constructed using exoplanets with both mass and radius measurements
(for example Weiss & Marcy 2014; Wolfgang et al. 2016; Chen & Kipping 2016) to
predict the expected masses of the transiting planets.
Table 7 lists gives constraints on the predicted masses of the exoplanet candidates
using the probabilistic M–R relationship from Chen & Kipping (2016) and taking
uncertainties in the planet radius and statistical uncertainties in the M–R parameters.
We compare the predictions with the predictions from Wolfgang et al. (2016) and find
similar values with smaller uncertainties and choose to use the Chen & Kipping (2016)
relationship because it is more conservative in the relevant range of parameter space.
A detailed discussion of the systematic effects introduced by the use of a M–R
relationship is beyond the scope of this paper but it is worth noting that all published
relationships are based on exoplanets much closer to their host star than any of the
candidates discussed here. This effect would cause the masses of these cool planets to
be systematically underestimated.
Using the same M–R relationship, we also compute the completeness of our transit
search as a function of planet mass and orbital semi-major axis. This function is
plotted in Figure 9 with the same color scheme as Figure 6. These injections and the
predicted masses and measured semi-major axes of the candidates can then be used
to estimate the occurrence rate in mass–semi-major axis units using the method from
Section 6. One small change to Equation (14) is necessary to account for the fact that
the injections were not made uniformly in ln M and ln a. We numerically estimate the
prior distribution in mass and semi-major axis from which the injections were drawn
p(ln a, ln M ) and Equation (14) becomes
Z(amin, amax; Mmin, Mmax)≈ K V
J
Qk(w(j))
p(ln a(j), ln M (j))
j=1
.
(18)
J(cid:88)
Using this result, we find that the mean occurrence rate density in the range 0.01 MJ ≤
M < 20 MJ and 1.5 au ≤ a < 9 au is
d2N
= 0.068 ± 0.019
d ln M d ln a
(19)
where N is the expected number of planets per G/K dwarf. This result and the
equivalent result as a function of planet mass and orbital period are listed in Table 8.
The uncertainty in Equation (19) and Table 8 does not take into account the uncer-
tainties in the mass estimates or any systematic noise in the mass–radius relationship.
Therefore, these specific results should be taken with the appropriate grain of salt but
predictions in these parameter spaces ease comparison with occurrence rates measured
computed using different methods.
The population of long-period transiting exoplanets
27
Clanton & Gaudi (2016) studied the occurrence rate of long-period giant planets
orbiting M dwarfs by combining results from radial velocity, microlensing, and direct
imaging surveys. In the period and mass range 103 − 104 days and 10 − 104 M⊕, they
find a mean occurrence rate density of
d2N
d ln M d ln P
= 0.023
(20)
per M dwarf with large uncertainty. This result is slightly lower than our estimated
rate for a similar range of masses and periods but around G/K dwarfs. This difference
is consistent with previous observational and theoretical results that cooler stars host
fewer long-period giant planets (for example Laughlin et al. 2004; Cumming et al.
2008; Clanton & Gaudi 2016).
Recently, Bryan et al. (2016) studied the frequency of long-period giant planets in
systems with inner hot Jupiters based on long-baseline radial velocity monitoring of
these systems (Knutson et al. 2014). In this sample, the computed occurrence rate of
long-period giant planets was found to be
d2N
= 0.125 ± 0.012
d ln M d ln a
(21)
in the range 1−20 MJ and 5−20 au. This result is about a factor of two larger than our
estimate (Equation 19) once again suggesting that cold Jupiters might preferentially
occur in systems with inner planets – or that the presence of cold Jupiters encourages
the formation of hot Jupiters.
A recent review of the occurrence rate estimates based on direct imaging surveys
(Bowler 2016) reports the upper limit on the occurrence rate of giant planets orbiting
F/G/K dwarfs as < 6.8% in the range 5 − 13 MJ and 10 − 100 au. Converted to a rate
density, this gives
d2N
d ln M d ln a
< 0.03 .
(22)
This value is lower than the value computed using our sample in Table 8 but this is
consistent with the fact that direct imaging is sensitive to the potentially less common
large planets at wider separations than detections with the transit method.
As a final comparison, we repeated the analysis of Burke et al. (2015) and fit a
double power-law occurrence rate to the short-period Kepler planet candidates7 and
extrapolated to the center of the two bins where we computed the occurrence rate. At
a period of 7 years and a radius of 0.2 RJ, the extrapolated occurrence rate density
is 0.73 ± 0.28 and at a radius of 0.6 RJ, the extrapolated rate density is 0.15 ± 0.05.
These extrapolated values are qualitatively consistent with the rates listed in Table 6
but we note that extrapolations and their statistical uncertainties should not be taken
too seriously.
7 The analysis was adapted from publicly available code that was demonstrated to reproduce the
same results as Burke et al. (2015) by Foreman-Mackey (2015).
28
Foreman-Mackey, Morton, Hogg, et al.
m
o
r
f
s
e
t
a
d
i
d
n
a
c
e
h
t
r
o
f
s
e
d
u
t
i
l
p
m
a
-
i
m
e
s
y
t
i
c
o
l
e
v
l
a
i
d
a
r
d
n
a
s
e
s
s
a
m
d
e
t
c
i
d
e
r
p
e
h
T
.
7
e
l
b
a
T
1
−
r
y
P
/
K
1
−
s
m
1
−
s
m
u
a
e
d
u
t
i
l
p
m
a
-
i
m
e
s
s
i
x
a
r
o
j
a
m
-
i
m
e
s
5
9
1
4
.
.
0
0
−
+
0
2
.
0
1
2
.
7
.
2
0
5
3
−
+
7
.
3
5
7
9
8
3
.
5
.
0
3
−
+
6
4
.
0
4
0
9
3
0
2
.
.
0
0
−
+
3
3
1
.
0
5
5
5
6
.
1
.
0
9
−
+
6
2
.
1
1
1
0
7
.
0
.
0
1
−
+
3
7
.
0
2
0
3
7
0
0
.
.
0
0
−
+
8
4
0
.
0
0
7
4
2
0
1
.
.
0
0
−
+
3
5
0
.
0
2
8
3
5
.
.
0
0
−
+
2
7
.
0
4
1
.
8
.
1
8
2
4
−
+
4
.
2
2
1
9
2
6
.
.
0
0
−
+
6
2
.
0
8
9
6
2
0
1
.
.
0
0
−
+
7
1
1
.
0
1
9
3
6
0
0
.
.
0
0
−
+
1
4
0
.
0
4
8
.
.
5
0
8
9
8
8
−
+
4
.
9
8
8
1
2
4
2
.
0
.
0
3
−
+
5
2
.
0
2
.
6
1
4
.
3
1
−
+
5
.
3
7
7
7
5
.
.
0
1
−
+
1
4
.
1
0
5
.
.
3
0
5
0
1
9
−
+
5
.
7
5
1
3
.
5
5
8
.
1
1
−
+
5
.
2
3
7
3
7
.
.
0
0
−
+
7
5
.
0
6
.
2
9
6
.
1
2
−
+
6
.
3
5
4
.
9
.
2
8
3
3
−
+
8
.
3
3
1
9
2
3
.
.
0
0
−
+
5
4
.
0
7
4
2
5
.
.
0
0
−
+
4
5
.
0
2
1
6
1
.
.
0
1
−
+
9
3
.
1
4
.
6
4
8
.
9
1
9
2
−
+
0
.
6
0
1
2
2
8
4
.
.
0
2
−
+
5
2
.
1
8
2
2
5
.
.
0
0
−
+
2
6
.
0
2
5
1
2
.
.
0
0
−
+
3
2
.
0
7
2
.
.
3
7
8
9
0
7
3
1
−
+
4
.
0
0
1
3
0
.
4
7
2
.
1
3
−
+
6
.
2
5
0
.
8
.
6
6
1
6
−
+
1
.
8
1
2
7
.
.
1
2
−
+
4
.
3
3
5
2
2
0
0
.
.
0
0
−
+
4
6
8
.
1
2
4
.
.
1
3
−
+
1
.
3
2
5
5
5
.
.
0
1
−
+
0
5
.
2
8
4
3
5
0
0
.
.
0
0
−
+
5
2
9
.
1
0
4
.
3
.
7
1
−
+
1
.
4
1
1
5
.
.
1
2
−
+
0
.
4
8
0
.
.
1
4
−
+
8
.
4
8
6
2
2
0
0
.
.
0
0
−
+
0
2
4
.
1
5
9
4
0
.
.
0
1
−
+
9
3
.
2
9
5
6
3
.
.
0
2
−
+
0
7
.
2
6
5
4
3
.
.
0
1
−
+
8
5
.
2
1
4
8
4
.
.
0
2
−
+
3
9
.
2
0
9
4
0
.
.
0
1
−
+
1
1
.
2
1
7
.
.
2
4
−
+
3
.
5
6
2
5
6
.
.
0
1
−
+
4
5
.
2
0
t
D
J
B
K
4
6
1
9
1
0
0
0
.
.
0
0
−
+
2
2
7
6
.
6
6
7
d
o
i
r
e
p
r
a
e
y
4
5
.
.
3
9
−
+
0
.
7
8
9
6
6
0
0
0
0
0
0
.
.
0
0
−
+
4
1
7
8
2
.
0
2
4
6
6
2
2
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
1
2
7
8
2
3
9
.
2
6
6
4
4
0
0
0
0
.
.
0
0
−
+
9
6
3
8
.
6
2
8
7
7
3
3
0
0
0
0
.
.
0
0
−
+
9
8
5
0
.
9
3
0
1
3
3
1
1
0
0
0
0
.
.
0
0
−
+
1
3
2
1
.
2
4
5
3
3
1
1
0
0
.
.
0
0
−
+
7
7
6
.
4
8
7
8
7
1
1
0
0
0
0
.
.
0
0
−
+
2
9
4
0
.
0
4
1
9
9
4
5
0
0
0
0
.
.
0
0
−
+
8
3
5
8
.
7
9
6
4
4
2
2
0
0
0
0
.
.
0
0
−
+
2
5
6
7
.
2
9
4
8
8
1
1
0
0
0
0
0
0
.
.
0
0
−
+
8
4
6
5
3
.
1
9
1
1
1
3
3
2
0
0
0
0
.
.
0
0
−
+
2
0
1
1
.
4
0
6
9
1
2
3
0
0
0
0
.
.
0
0
−
+
6
7
9
5
.
3
9
3
3
4
6
6
0
0
0
0
.
.
0
0
−
+
2
6
5
3
.
4
5
5
5
5
1
1
0
0
0
0
0
0
.
.
0
0
−
+
2
9
8
0
8
.
0
3
8
7
7
4
4
0
0
0
0
.
.
0
0
−
+
4
4
3
2
.
6
2
2
6
8
1
1
0
0
0
0
.
.
0
0
−
+
4
7
6
2
.
7
5
6
8
0
.
1
.
3
1
−
+
9
.
5
2
2
.
.
1
4
−
+
0
.
4
4
3
5
5
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
7
9
0
8
8
6
8
.
2
1
4
.
.
6
8
3
8
−
+
2
.
4
5
4
5
.
.
3
9
−
+
1
.
9
9
0
.
4
.
5
1
−
+
9
.
9
1
2
9
9
0
0
0
0
0
0
0
0
0
0
.
.
0
0
−
+
7
5
9
9
7
2
9
.
1
1
3
.
.
1
3
−
+
3
.
4
8
6
.
.
1
7
−
+
9
.
4
3
3
.
.
1
4
−
+
9
.
4
1
1
.
.
2
8
−
+
5
.
5
3
5
8
6
.
.
0
2
−
+
6
1
.
3
2
6
.
0
.
6
2
−
+
7
.
2
1
3
7
.
.
1
4
−
+
3
.
4
s
s
a
m
J
M
8
4
3
7
0
0
.
.
0
0
−
+
9
7
0
.
0
1
3
.
8
.
6
3
−
+
5
.
6
8
9
7
8
0
2
.
.
0
9
−
+
0
4
1
.
0
5
3
1
3
0
0
.
.
0
0
−
+
8
2
0
.
0
2
1
8
8
3
0
.
1
.
0
1
−
+
7
5
1
.
0
5
6
.
7
.
3
4
−
+
8
.
3
2
2
1
2
0
0
.
.
0
0
−
+
8
2
0
.
0
9
7
1
3
0
0
.
.
0
0
−
+
2
4
0
.
0
2
9
2
3
0
0
.
.
0
0
−
+
9
4
0
.
0
9
3
.
7
.
4
9
−
+
6
.
4
6
6
3
9
0
0
.
.
0
0
−
+
1
6
0
.
0
2
2
1
2
0
0
.
.
0
0
−
+
6
2
0
.
0
8
5
5
1
0
1
0
0
.
.
0
0
−
+
0
2
1
0
.
0
6
.
9
1
.
6
3
1
7
−
+
4
.
2
6
1
5
2
1
1
.
2
.
0
2
−
+
9
1
.
0
4
2
4
8
.
5
.
0
3
−
+
3
9
.
0
2
e
l
b
a
T
i
s
u
d
a
r
p
K
d
i
c
i
k
J
R
3
2
9
9
0
0
.
.
0
0
−
+
4
1
5
.
0
9
9
3
3
0
0
.
.
0
0
−
+
6
7
8
.
0
5
6
1
1
.
.
0
0
−
+
7
6
.
0
3
3
8
9
0
0
.
.
0
0
−
+
2
8
2
.
0
8
7
7
0
0
1
.
.
0
0
−
+
8
9
6
.
0
5
0
2
3
.
.
0
0
−
+
4
0
.
1
7
7
1
1
0
0
.
.
0
0
−
+
7
7
2
.
0
4
5
4
4
0
0
.
.
0
0
−
+
5
5
3
.
0
5
5
2
2
0
0
.
.
0
0
−
+
6
8
3
.
0
6
9
3
4
.
.
0
0
−
+
2
2
.
1
3
1
1
2
.
.
0
0
−
+
3
4
.
0
4
7
2
2
0
0
.
.
0
0
−
+
6
6
2
.
0
7
6
3
4
0
0
.
.
0
0
−
+
3
6
1
.
0
5
6
3
6
.
.
0
0
−
+
0
0
.
2
6
6
1
1
.
.
0
0
−
+
4
7
.
0
1
2
1
1
.
.
0
0
−
+
3
8
.
0
6
.
4
1
0
.
4
1
9
.
4
1
6
.
3
1
4
.
3
1
6
.
3
1
0
.
3
1
9
.
3
1
4
.
3
1
0
.
4
1
4
.
4
1
4
.
3
1
9
.
1
1
9
.
4
1
6
.
4
1
5
.
4
1
8
0
9
8
1
2
3
5
4
9
9
3
2
3
0
6
4
4
5
7
4
0
4
4
1
5
5
6
7
9
6
0
1
4
8
7
5
9
6
2
4
8
5
1
2
5
0
5
8
5
3
7
8
3
7
8
4
5
9
0
0
8
8
7
0
3
6
0
3
9
9
5
1
7
8
1
0
1
3
2
7
7
8
2
0
1
9
1
3
1
2
3
0
1
8
6
0
2
0
6
0
1
8
1
7
2
4
8
0
1
4
2
1
9
0
7
1
1
n
e
h
C
(
n
o
i
t
a
l
e
r
i
s
u
d
a
r
–
s
s
a
m
c
i
t
s
i
l
i
b
a
b
o
r
p
a
g
n
i
s
u
i
s
u
d
a
r
d
e
r
u
s
a
e
m
e
h
t
n
o
d
e
s
a
b
d
e
t
a
m
i
t
s
e
e
r
a
s
e
d
u
t
i
l
p
m
a
y
t
i
c
o
l
e
v
l
a
i
d
a
r
d
n
a
s
e
s
s
a
m
e
h
T
-
e
t
o
N
.
)
6
1
0
2
i
g
n
p
p
K
&
i
The population of long-period transiting exoplanets
29
Table 8. The occurrence rate computed in mass units
volume
2 yr < P < 25 yr; 0.01 MJ ≤ M < 20 MJ
1.5 au < a < 9 au; 0.01 MJ ≤ M < 20 MJ
aThe rate density is measured in natural logarithmic units; see Equation (19).
Note-These values are computed assuming that the occurrence rate is flat
rate densitya
0.046 ± 0.013
0.068 ± 0.019
integrated rate
0.882 ± 0.245
0.925 ± 0.257
in the logarithmic parameters.
9. PROSPECTS FOR FOLLOW-UP
A real concern about the detection of exoplanets from a single transit is that follow-
up and confirmation is difficult. Since the period of the orbit is poorly constrained and
transits are sparse, any prediction of a subsequent transit time will be too uncertain
to schedule targeted photometric follow-up (Beichman et al. 2016; Dalba & Muirhead
2016). Instead, follow-up using radial velocity and astrometry are more promising. For
both radial velocity and astrometry, there is information about the orbiting planets in
measurements made at all times – not just during transit. This allows observations to
be scheduled without a well constrained orbital period. Furthermore, follow-up of any
of these candidates using radial velocity or astrometry would provide a measurement of
the density of a planet that would be valuable for the study of planetary compositions.
Table 7 lists the posterior predictions for the semi-amplitude K of the radial velocity
signal produced by each candidate using the mass predictions from the previous section.
Since the orbital periods are long, we also include a simple prediction for the slope
of the radial velocity trend induced by this planet by taking the ratio of the semi-
amplitude and the orbital period. Any radial velocity follow-up of the candidates
presented here would be an ambitious undertaking because the stars are relatively
faint and, in most cases, the radial velocity trends are small. Some candidates should,
however, be within the reach of current state-of-the-art facilities.
In principle the Gaia Mission will be very sensitive to the astrometric wobble
produced by a long-period exoplanet (Perryman et al. 2014). To leading order,
the astrometric signal strength is proportional to the semi-major axis of the stellar
(primary) reflex motion in angular units. That is, detectability is related to the angle
α given by
α =
a
D
Mp
Ms
,
(23)
where a is the semi-major axis, D is the distance from the observer to the primary,
and Mp/Ms is the planet-to-star mass ratio.
The single-visit precision of Gaia will vary with magnitude but is expected to
be on the order of 40 µas at these magnitudes. In detail the confidence with which
an exoplanet can be detected or measured in the final Gaia data depends on this
30
Foreman-Mackey, Morton, Hogg, et al.
Figure 9. The same as Figure 6 converted into the planet mass and semi-major axis
plane. Since the completeness function depends on the planet's radius and not its mass, a
probabilistic mass–radius relationship (Chen & Kipping 2016) was used to convert radius to
mass.
248semi-majoraxis[au]0.010.1110100MP/MJ0.1140.5250.7070.7110.6980.6330.5780.1080.4880.6500.6360.6230.5650.5150.1050.4760.6180.5920.5760.5170.4740.0890.4330.5290.5000.4810.4230.3930.00.30.60.00.30.6The population of long-period transiting exoplanets
31
precision, the number of crossings of the star through the Gaia field-of-view, and
details of how the projected orbit is sampled by the time history of the focal-plane
crossings. However, it is not expected that Gaia can detect or precisely measure
exoplanet-induced astrometric wobbles that are much smaller in amplitude than the
single-visit precision (Perryman et al. 2014).
The primary stars in the Kepler Field are typically at distances of ∼ 500 pc, and
typical mass ratios are in the 10−4 range. We therefore expect astrometric amplitudes
in the 0.3 to 3 µas range. These planets will not be detectable or measurable in
the Gaia data under any circumstances, but it may be possible to identify which
candidates are false-alarms caused by eclipsing binaries. However, similar planets
around closer stars will be detectable with Gaia. This means that there will be a
comparable exoplanet occurrence rate measurement from the Gaia data. It also means
that many of the discoveries of the K2 and TESS missions could be followed up and
precisely measured by the Gaia Mission.
10. SUMMARY
We have developed a fully automated method to search for the transits of long-
period exoplanets with only one or two observable transits in the Kepler archival light
curves. This method uses probabilistic model comparison to veto non-transit signals.
Applying this method to the brightest 39,036 G/K dwarfs in the Kepler target list,
we discover 16 systems with likely astrophysical transits and eclipses. We fit the light
curve for each candidate with a physical generative model and informative priors on
eccentricity and stellar density to estimate the planet's orbital period. The constraint
on the period is also informed by the simplifying assumption that no other transit
could occur during the baseline of Kepler observations of the target. Simulations of
the false positive population – lone primary or secondary eclipses of binary systems
or background eclipsing binaries – suggest that 13 of these candidates have high
probability of being planetary in nature.
We measure the empirical detection efficiency function of our search procedure
by injecting simulated transit signals into the target light curves and measuring
the recovery rate. By combining the measured detection efficiency and the catalog
of exoplanet candidates, we estimate the integrated occurrence rate of exoplanets
with orbital periods in the range 2 − 25 years and radii in the range 0.1 − 1 RJ to
be 2.00 ± 0.72 planets per G/K dwarf. This result is qualitatively consistent with
estimates of the occurrence rate of long-period giant planets based on data from radial
velocity and direct imaging surveys. The occurrence rate measured here – for Sun-like
hosts – is higher than microlensing results for generally lower mass stars (Gaudi 2012;
Clanton & Gaudi 2014, 2016) but this discrepancy is consistent with predictions from
the core-accretion model (Laughlin et al. 2004).
Using a probabilistic mass–radius relationship, we predict the masses of our candi-
dates and report predictions for the radial velocity semi-amplitudes. Unfortunately,
32
Foreman-Mackey, Morton, Hogg, et al.
since the target stars are faint and the amplitudes are small, these targets are unlikely
to be accessible with even the current state-of-the-art high-precision instruments. We
also discuss the potential for astrometric follow-up using the forthcoming data from
the Gaia Mission with similarly discouraging results.
Any detailed analysis of individual systems detected with only a single transit
requires follow-up observations to convincingly rule out false positive scenarios and to
better characterize the stellar host parameters (with, for example, parallax measure-
ments from Gaia). The conclusions of this work – and all other occurrence rate results
based on Kepler data – are conditioned on the assumption that the stellar characteri-
zation of the target sample is systematic and un-biased. The main population-level
results should be fairly insensitive systematic issues with the sample but a rigorous
analysis of these effects will be required to come to more detailed conclusions about
this population of long-period transiting planets.
Our method of transit discovery is especially relevant for future photometric surveys
like K2, TESS, and PLATO where the survey baseline is shorter than Kepler. The
transits of planets with orbital periods longer than the observation baselines will be
plentiful in these forthcoming data sets and this method can, in principle, be trivially
generalized to discover these planets, prioritize follow-up, and study their population.
All of the code used in this project is available from https://github.com/dfm/
peerless under the MIT open-source software license. This code (plus some depen-
dencies) can be run to re-generate all of the figures and results in this paper; this
version of the paper was generated with git commit a485057 (2016-09-30). The param-
eter estimation results represented as MCMC samplings and the injection results are
available for download from Zenodo at http://dx.doi.org/10.5281/zenodo.58273.
It is a pleasure to thank Jeff Coughlin, So Hattori, Heather Knutson, Phil Muirhead,
Darin Ragozzine, Hans-Walter Rix, Dun Wang, and Angie Wolfgang for helpful
discussions and contributions. We thank the anonymous referee for comments that
improved the presentation and clarity of this manuscript.
T.D.M. was supported by the National Aeronautics and Space Administration,
under the Kepler participating scientist program (grant NNX14AE11G), and is grateful
for the hospitality of both the Institute for Advanced Study and Carnegie Observatories
that helped support this work. D.W.H. was partially supported by the National Science
Foundation (grant IIS-1124794), the National Aeronautics and Space Administration
(grant NNX12AI50G), and the Moore–Sloan Data Science Environment at NYU. E.A.
acknowledges support from NASA grants NNX13AF20G, NNX13AF62G, and NASA
Astrobiology Institutes Virtual Planetary Laboratory, supported by NASA under
cooperative agreement NNH05ZDA001C.
The population of long-period transiting exoplanets
33
This research made use of the NASA Astrophysics Data System and the NASA
Exoplanet Archive. The Exoplanet Archive is operated by the California Institute of
Technology, under contract with NASA under the Exoplanet Exploration Program.
This paper includes data collected by the Kepler mission. Funding for the Kepler
mission is provided by the NASA Science Mission directorate. We are grateful to the
entire Kepler team, past and present. Their tireless efforts were all essential to the
tremendous success of the mission and the successes of K2, present and future.
These data were obtained from the Mikulski Archive for Space Telescopes (MAST).
STScI is operated by the Association of Universities for Research in Astronomy, Inc.,
under NASA contract NAS5-26555. Support for MAST is provided by the NASA
Office of Space Science via grant NNX13AC07G and by other grants and contracts.
Computing resources were provided by High Performance Computing at New York
University.
Facility: Kepler
Software: batman (Kreidberg 2015), ceres (Agarwal et al. 2016), corner.py
(Foreman-Mackey 2016), emcee (Foreman-Mackey et al. 2013), exosyspop (Morton &
Foreman-Mackey 2016), george (Ambikasaran et al. 2016), isochrones (Morton 2015a),
matplotlib (Hunter et al. 2007), numpy (Van Der Walt et al. 2011), scipy (Jones et al.
2001), transit (Foreman-Mackey & Morton 2016), vespa (Morton 2015b).
34
Foreman-Mackey, Morton, Hogg, et al.
APPENDIX
A. DETAILS OF THE LIGHT CURVE MODELS
In Section 2.2, the five light curve models were listed. In this section, we give
the mathematical details of each model and list the parameters that are fit. Each
model – except the transit model – can be easily differentiated with respect to its
parameters. As discussed in the following section, this feature is crucial for efficient
and robust likelihood maximization.
• The box model is given by
mbox(t) =
a, if t ≤ tmin
b, if tmin < t ≤ tmax
c, if tmax < t
(A1)
(A2)
(A3)
where a, b, and c are free parameters, and tmin and tmax are fixed. In practice,
we include two different box models where tmin and tmax are set using different
heuristics. The first box has the bounds set to match the ingress and egress of
the best fit transit. The second box is chosen based on the largest change points
in the light curve.
• The step model is given by
m1 + h1 exp ([t − t0]/w1) , if t < t0
m2 + h2 exp ([t0 − t]/w2) , if t0 ≤ t
mstep(t) =
where all of the parameters – including t0 – are included in the fit. To ensure
that the widths w1 and w2 remain positive, we fit for log w1 and log w2.
• For the outlier model, we iterate through all cadences tn within 0.3 days of the
candidate transit time and evaluate the model as
f (t0),
moutlier(t) =
if t = t0
median[f (t (cid:54)= t0)], if t (cid:54)= t0
where f (t) is the observed time series. With this model, no non-linear optimization
is required and the final value of t0 is the one with the maximum likelihood in
this grid search.
• The variability model only has one parameter, the flux m0 and mvariability(t) = m0
at all times. The variability is captured by the Gaussian Process residual model.
• Finally, the transit model is an exposure time integrated, limb darkened light curve
(Mandel & Agol 2002; Kipping 2010) parameterized by the radius ratio between
the planet and star, the transit duration, the transit time, the impact parameter,
The population of long-period transiting exoplanets
35
and two quadratic limb darkening coefficients (Kipping 2013a). Analytically
computing the gradient of a simple transit model is possible (P´al 2008) but it
becomes substantially more tedious as the model becomes more realistic. There-
fore, we instead use a compile-time automatic differentiation library8 (Agarwal
et al. 2016) to efficiently compute first derivatives of the full transit model with
respect to the orbital and physical parameters to machine precision.
B. GAUSSIAN PROCESS REGRESSION
Gaussian Processes (GPs) are a class of non-parametric, stochastic models that
have been demonstrated to be good effective models for the variability in Kepler light
curves. A simple GP model can be used to capture residual non-transit variability in
light curves. In this paper, we use a GP model for two steps: light curve–level transit
shape vetting and parameter estimation. A full discussion of GPs is beyond the scope
of this paper, so we will only summarize the most relevant points here and direct an
interested reader to Rasmussen & Williams (2006) for more details.
A GP model is specified by the following likelihood function
L = ln p(y θ, α) =−1
2
r(θ)T K(α)−1 r(θ) − 1
2
log det K(α) − N
2
log 2 π (B4)
where y is a list of measurements in a scalar time series – in this case, fluxes – measured
at the times t, and
r(θ) = y − m(t; θ)
is the vector of residuals away from the mean model m(t; θ). For the purposes of this
paper, we model the covariance matrix K(α) using the Mat´ern-3/2 kernel. Under this
model, the elements of K(α) are given by
[K(α)]ij = σ2
i δij + α2
1 +
(cid:20)
(cid:21)
(cid:18)
ti − tj√
3 τ
exp
−ti − tj√
3 τ
(cid:19)
(B5)
(B6)
where σi is the reported uncertainty on the i-th measurement in the time series and
δij is the Kronecker delta.
This covariance function (Equation B6) is specified by an amplitude α and a time
scale τ and we will simultaneously fit for these hyperparameters α = (α, τ ) and the
parameters of the mean model θ. To efficiently find the parameter set that maximizes
Equation (B4) using a non-linear optimization routine9, it is useful to be able to
compute the gradient of Equation (B4) with respect to the parameters θ and α. These
gradients are given by
d ln p(y θ, α)
dθ
=
dm(t; θ)
T
dθ
K(α)−1 r(θ)
(B7)
8 More specifically, we use the Jet object from the BSD-licensed Ceres Solver http://
ceres-solver.org
9 We use the L-BFGS-B method as implemented in SciPy http://docs.scipy.org/doc/scipy/
reference/generated/scipy.optimize.minimize.html.
36
and
where
Foreman-Mackey, Morton, Hogg, et al.
(cid:18)(cid:2)φ φT − K(α)−1(cid:3) dK(α)
dα
d ln p(y θ, α)
dα
=
1
2
Tr
φ = K(α)−1 r(θ)
.
REFERENCES
(cid:19)
(B8)
(B9)
Agarwal, S., Mierle, K., & et al. 2016, Ceres
Foreman-Mackey, D. 2015, An experiment in
Solver, http://ceres-solver.org, ,
Ambikasaran, S., Foreman-Mackey, D.,
Greengard, L., Hogg, D., & O Neil, M. 2016,
IEEE
Ballard, S., & Johnson, J. A. 2016, ApJ, 816,
66
open science: exoplanet population inference,
http://dfm.io/posts/exopop/, Zenodo,
doi:10.5281/zenodo.57218
-. 2016, The Journal of Open Source
Software, 24, doi:10.21105/joss.00024
Foreman-Mackey, D., Hogg, D. W., Lang, D.,
Beichman, C., Livingston, J., Werner, M., et al.
& Goodman, J. 2013, PASP, 125, 306
2016, ApJ, 822, 39
Foreman-Mackey, D., Hogg, D. W., & Morton,
Borucki, W. J., Koch, D. G., Basri, G., et al.
T. D. 2014, ApJ, 795, 64
2011, ApJ, 736, 19
Bowler, B. P. 2016, ArXiv e-prints,
arXiv:1605.02731
Bryan, M. L., Knutson, H. A., Howard, A. W.,
et al. 2016, ApJ, 821, 89
Bryson, S. T., Jenkins, J. M., Gilliland, R. L.,
et al. 2013, PASP, 125, 889
Burke, C. J., & McCullough, P. R. 2014, ApJ,
792, 79
Burke, C. J., Bryson, S. T., Mullally, F., et al.
2014, ApJS, 210, 19
Foreman-Mackey, D., Montet, B. T., Hogg,
D. W., et al. 2015, ApJ, 806, 215
Foreman-Mackey, D., & Morton, T. 2016,
dfm/transit: v0.3.0, , ,
doi:10.5281/zenodo.159478
Gaidos, E., & Mann, A. W. 2012, ApJ, 762, 41
Gaudi, B. S. 2012, ARA&A, 50, 411
Girardi, L., Groenewegen, M. A. T.,
Hatziminaoglou, E., & da Costa, L. 2005,
A&A, 436, 895
Gould, A., Gaudi, B. S., & Han, C. 2004,
Burke, C. J., Christiansen, J. L., Mullally, F.,
et al. 2015, ApJ, 809, 8
ArXiv Astrophysics e-prints,
astro-ph/0405217
Cassan, A., Kubas, D., Beaulieu, J.-P., et al.
Gould, A., Dong, S., Gaudi, B. S., et al. 2010,
2012, Nature, 481, 167
ApJ, 720, 1073
Chen, J., & Kipping, D. M. 2016, ArXiv
Howard, A. W., Marcy, G. W., Bryson, S. T.,
e-prints, arXiv:1603.08614
et al. 2012, ApJS, 201, 15
Christiansen, J. L., Clarke, B. D., Burke, C. J.,
Howell, S. B., Sobeck, C., Haas, M., et al.
et al. 2013, ApJS, 207, 35
-. 2015, ApJ, 810, 95
Clanton, C., & Gaudi, B. S. 2014, ApJ, 791, 91
-. 2016, ApJ, 819, 125
Coughlin, J. L., Mullally, F., Thompson, S. E.,
et al. 2016, ApJS, 224, 12
Cumming, A., Butler, R. P., Marcy, G. W.,
et al. 2008, PASP, 120, 531
Dalba, P. A., & Muirhead, P. S. 2016, ApJL,
826, L7
Dalba, P. A., Muirhead, P. S., Fortney, J. J.,
et al. 2015, ApJ, 814, 154
2014, PASP, 126, 398
Huber, D., Silva Aguirre, V., Matthews, J. M.,
et al. 2014, ApJS, 211, 2
Hunter, J. D., et al. 2007, Computing in
science and engineering, 9, 90
Jones, E., Oliphant, T., Peterson, P., et al.
2001, SciPy: Open source scientific tools for
Python, ,
Kipping, D. M. 2010, MNRAS, 408, 1758
-. 2013a, MNRAS, 435, 2152
-. 2013b, MNRAS, 434, L51
Kipping, D. M., Bastien, F. A., Stassun, K. G.,
Dawson, R. I., & Johnson, J. A. 2012, ApJ,
et al. 2014a, ApJL, 785, L32
756, 122
Dressing, C. D., & Charbonneau, D. 2015,
ApJ, 807, 45
Fang, J., & Margot, J.-L. 2012, ApJ, 761, 92
Fischer, D. A., Schwamb, M. E., Schawinski,
K., et al. 2012, MNRAS, 419, 2900
Kipping, D. M., Dunn, W. R., Jasinski, J. M.,
& Manthri, V. P. 2012, MNRAS, 421, 1166
Kipping, D. M., Torres, G., Buchhave, L. A.,
et al. 2014b, ApJ, 795, 25
Kipping, D. M., Torres, G., Henze, C., et al.
2016, ApJ, 820, 112
The population of long-period transiting exoplanets
37
Kirk, B., Conroy, K., Prsa, A., et al. 2016, AJ,
Rasmussen, C. E., & Williams, C. K. I. 2006,
151, 68
Knutson, H. A., Fulton, B. J., Montet, B. T.,
et al. 2014, ApJ, 785, 126
Kov´acs, G., Zucker, S., & Mazeh, T. 2002,
A&A, 391, 369
Kreidberg, L. 2015, PASP, 127, 1161
Laughlin, G., Bodenheimer, P., & Adams,
F. C. 2004, ApJL, 612, L73
Mandel, K., & Agol, E. 2002, ApJL, 580, L171
Moriarty, J., & Ballard, S. 2015, ArXiv
e-prints, arXiv:1512.03445
Morton, T., & Foreman-Mackey, D. 2016,
timothydmorton/exosyspop: v0.1-alpha, , ,
doi:10.5281/zenodo.159487
Morton, T. D. 2012, ApJ, 761, 6
-. 2015a, isochrones: Stellar model grid
package, Astrophysics Source Code Library, ,
, ascl:1503.010
-. 2015b, VESPA: False positive probabilities
calculator, Astrophysics Source Code
Library, , , ascl:1503.011
Morton, T. D., Bryson, S. T., Coughlin, J. L.,
et al. 2016, ApJ, 822, 86
Morton, T. D., & Johnson, J. A. 2011, ApJ,
738, 170
Mulders, G. D., Pascucci, I., & Apai, D. 2015,
ApJ, 814, 130
Mullally, F., Coughlin, J. L., Thompson, S. E.,
et al. 2016, PASP, 128, 074502
Osborn, H. P., Armstrong, D. J., Brown,
D. J. A., et al. 2016, MNRAS, 457, 2273
P´al, A. 2008, MNRAS, 390, 281
Perryman, M., Hartman, J., Bakos, G. ´A., &
Lindegren, L. 2014, ApJ, 797, 14
Petigura, E. A., Howard, A. W., & Marcy,
G. W. 2013, Proceedings of the National
Academy of Science, 110, 19273
Gaussian processes for machine learning
(MIT Press)
Rauer, H., Catala, C., Aerts, C., et al. 2014,
Experimental Astronomy, 38, 249
Ricker, G. R., Winn, J. N., Vanderspek, R.,
et al. 2015, Journal of Astronomical
Telescopes, Instruments, and Systems, 1,
014003
Rogers, L. A. 2015, ApJ, 801, 41
Rowe, J. F., Coughlin, J. L., Antoci, V., et al.
2015, ApJS, 217, 16
Shvartzvald, Y., Maoz, D., Udalski, A., et al.
2016, MNRAS, 457, 4089
Smith, J. C., Stumpe, M. C., Van Cleve, J. E.,
et al. 2012, PASP, 124, 1000
Stumpe, M. C., Smith, J. C., Van Cleve, J. E.,
et al. 2012, PASP, 124, 985
Tremaine, S., & Dong, S. 2012, AJ, 143, 94
Uehara, S., Kawahara, H., Masuda, K.,
Yamada, S., & Aizawa, M. 2016, ApJ, 822, 2
Van Der Walt, S., Colbert, S. C., &
Varoquaux, G. 2011, Computing in Science
& Engineering, 13, 22
Wang, J., Fischer, D. A., Barclay, T., et al.
2013, ApJ, 776, 10
-. 2015, ApJ, 815, 127
Weiss, L. M., & Marcy, G. W. 2014, ApJL,
783, L6
Winn, J. N. 2010, ArXiv e-prints,
arXiv:1001.2010
Wolfgang, A., Rogers, L. A., & Ford, E. B.
2016, ApJ, 825, 19
Yee, J. C., & Gaudi, B. S. 2008, ApJ, 688, 616
|
1612.02231 | 1 | 1612 | 2016-12-07T12:52:57 | The temporal evolution of exposed water ice-rich areas on the surface of 67P/Churyumov-Gerasimenko: spectral analysis | [
"astro-ph.EP"
] | Water ice-rich patches have been detected on the surface of comet 67P/Churyumov-Gerasimenko by the VIRTIS hyperspectral imager on-board the Rosetta spacecraft, since the orbital insertion in late August 2014. Among those, three icy patches have been selected, and VIRTIS data are used to analyse their properties and their temporal evolution while the comet was moving towards the Sun. We performed an extensive analysis of the spectral parameters, and we applied the Hapke radiative transfer model to retrieve the abundance and grain size of water ice, as well as the mixing modalities of water ice and dark terrains on the three selected water ice rich areas. Study of the spatial distribution of the spectral parameters within the ice-rich patches has revealed that water ice follows different patterns associated to a bimodal distribution of the grains: ~50 {\mu}m sized and ~2000 {\mu}m sized. In all three cases, after the first detections at about 3.5 AU heliocentric distance, the spatial extension and intensity of the water ice spectral features increased, it reached a maximum after 60-100 days at about 3.0 AU, and was followed by an approximately equally timed decrease and disappearanceat about ~2.2 AU, before perihelion. The behaviour of the analysed patches can be assimilated to a seasonal cycle. In addition we found evidence of short-term variability associated to a diurnal water cycle. The similar lifecycle of the three icy regions indicates that water ice is uniformly distributed in the subsurface layers, and no large water ice reservoirs are present. | astro-ph.EP | astro-ph | The temporal evolution of exposed water ice-rich areas on the surface of
67P/Churyumov-Gerasimenko: spectral analysis.
A. Raponi1, M. Ciarniello1, F. Capaccioni1, G. Filacchione1, F. Tosi1, M. C. De Sanctis1, M.T. Capria1, M. A.
Barucci2, A. Longobardo1, E. Palomba1, D. Kappel3, G. Arnold3, S. Mottola3, B. Rousseau2, E. Quirico4, G.
Rinaldi1, S. Erard2, D. Bockelee-Morvan2, C. Leyrat2
1IAPS-INAF, via del Fosso del Cavaliere 100, Rome 00133, Italy
2LESIA, Observatoire de Paris/CNRS/Université Pierre et Marie Curie/Université Paris-Diderot, Meudon, France
3Institute for Planetary Research, Deutsches Zentrum fur Luft- und Raumfahrt (DLR), Berlin, Germany
4Université Grenoble Alpes, CNRS, IPAG, Grenoble, France
Abstract. Water ice-rich patches have been detected on the surface of comet 67P/Churyumov-Gerasimenko
by the VIRTIS hyperspectral imager on-board the Rosetta spacecraft, since the orbital insertion in late
August 2014. Among those, three icy patches have been selected, and VIRTIS data are used to analyse their
properties and their temporal evolution while the comet was moving towards the Sun. We performed an
extensive analysis of the spectral parameters, and we applied the Hapke radiative transfer model to retrieve
the abundance and grain size of water ice, as well as the mixing modalities of water ice and dark terrains on
the three selected water ice rich areas.
Study of the spatial distribution of the spectral parameters within the ice-rich patches has revealed that
water ice follows different patterns associated to a bimodal distribution of the grains: ~50 μm sized and
~2000 μm sized.
In all three cases, after the first detections at about 3.5 AU heliocentric distance, the spatial extension and
intensity of the water ice spectral features increased, it reached a maximum after 60-100 days at about 3.0
AU, and was followed by an approximately equally timed decrease and disappearance at about ~2.2 AU,
before perihelion. The behaviour of the analysed patches can be assimilated to a seasonal cycle. In addition
we found evidence of short-term variability associated to a diurnal water cycle.
The similar lifecycle of the three icy regions indicates that water ice is uniformly distributed in the
subsurface layers, and no large water ice reservoirs are present.
Introduction
Comets are primordial bodies, which played a key role in the early phases of solar system
formation contributing to the accretion of the giant planets as well as potentially contributing
volatiles to the terrestrial planets (A'Hearn et al., 2011a). Water is the most abundant volatile
species in comets as demonstrated by long-term studies of cometary comae (e.g. Bockelée-
Morvan and Rickman, 1997; Bockelée-Morvan et al., 2004; Feaga et al., 2007; Mumma and
Charnley, 2011).
Water was assembled, presumably, as amorphous ice within the cometesimal, as a consequence of
the very low formation temperatures of the icy grains. Evolution of the surface, and of the ice
therein contained, of Jupiter Family Comets (JFC) is driven by their periodic passages close to the
Sun. The surface is modified by sublimation of volatiles leading to surface erosion, which causes
material's loss up to tens of meters at each perihelion passage, and by heating, which alters the
structure of the upper layers of the nucleus to a depth determined by the thermal skin depth. In
summary, at each periodic passage within the inner Solar System, primordial water ice is subjected
to evolutionary processes such as phase transition, sublimation, re-condensation, sintering.
Amorphous ice is likely not observable directly at the surface of a cometary nucleus at close
distances to the Sun as it undergoes a phase transition with an activation temperature of 120 K,
which transforms amorphous ice into crystalline ice.
The study of the state of water ice (crystalline Vs. amorphous, grain size, presence of clathrates,
etc.) at the surface of a cometary nucleus can provide some insight into the processes that have
led to its evolution.
The properties investigated by the present work are: abundance, grain size, and mixing modalities
of the ice with the refractory dark material of comet. Hereafter, the abundance is intended as the
total cross section of the icy grains as a fraction of the area subtended by the pixel. The grain size is
the physical diameter of the spherical grain, as in Hapke (2012). The mixing modalities are
assumed to be either intimate or areal. In the intimate mixing the grains of the ice and dark terrain
are in contact each other. In the areal mixing the ice is present in form of patches on the surface.
A first insight on the properties cometary water ice was provided by the in situ observations
carried out by the Deep Impact mission to 9P/Tempel 1 (A'Hearn et al., 2005) and its investigation
of 103P/Hartley 2 during the extended mission (A'Hearn et al., 2011b). However, a real
breakthrough was represented by the Rosetta mission to the comet 67P/Churyumov-Gerasimenko
(hereafter referred to as 67P) which carried out the most complete set of observations, performing
the first escort and soft landing on a comet.
The more diffuse form in which the water ice has been detected in different comets is a population
of pure and very fine grains (~1 μm in diameter) of crystalline ice:
- Ground observation of the large and infrequent outbursts of comet 17P/Holmes revealed pure
and fine water ice grains (Yang et al., 2009);
- The material excavated and ejected from 10 to 20-m depths from comet 9P/Tempel 1 by the
impactor released by the Deep Impact spacecraft has been modelled with 1 μm pure water ice
grains (Sunshine et al., 2007);
- The innermost coma of the hyperactive comet 103P/Hartley 2 has been found to be populated by
1 μm sized grains, which are components of most massive aggregates dragged out continuously by
gaseous CO2 (Protopapa et al., 2014);
- Observations, by the VIRTIS instrument (Coradini et al., 2007) on-board the Rosetta spacecraft, of
the neck region on the surface of 67P revealed fine grains (1-2 μm) intimately mixed with the
predominant dark terrain, which has been suggested to be the result of a diurnal sublimation-
recondensation cycle (De Sanctis et al., 2015);
- An increasing exposure of fine grains (1-2 μm) of water ice on 67P observed after the gradual
removal of the surficial dust layer by the gaseous activity, before perihelion (Filacchione et al.,
2016b; Ciarniello et al., submitted).
However, there is at least another form of cometary water ice: in some specific and small areas on
the surface of comets it has been observed in form of patches with coarser grain sizes (> 30 μm).
Specifically:
- On the surface of comet Tempel 1 water ice has been detected by the HRI spectrometer onboard
the Deep Impact spacecraft in three regions covering a total area of 0.5 km2. The icy areas have
been modeled with a few percent of water ice constituted by 30 μm sized grains, in areal mixture
with the dark terrain of the comet (Sunshine et al., 2006), and by up to 1%, with 30 or 70 μm sized
grains, in areal or intimate mixture, respectively (Raponi et al., 2013);
- On the southern hemisphere of 67P icy patches have been detected in correspondence of two
debris falls. Spectral modelling has revealed a bimodal distribution of grains: ~50 μm sized and
~2000 μm sized, respectively in intimate and areal mixture with the dark refractory material of the
comet (Filacchione et al., 2016a);
- Other dozen small patches or cluster of spots spread over the entire surface of 67P, and covering
up to few hundred square meters each one, have been detected by OSIRIS (Keller et al., 2007) and
VIRTIS instruments on board the Rosetta spacecraft (Pommerol et al., 2015, Barucci et al., 2016).
In order to have an overall view of the physical state and processes of water ice on comets,
intrinsic differences between the comets observed, in terms of composition, activity and surface
morphology must be taken into account. The evolution of comets is then subjected to seasonal
effects, related to the total solar input received by the comets along an entire orbit, and to short
term processes, attributable to the instantaneous solar input.
The analysis of the continuous evolution in time of the water ice on the surface of comets was not
possible before the Rosetta mission, because all the previous space missions performed scientific
measurements during only relatively short flybys. It is worth noting that one example existed,
before Rosetta, of a long term analysis of a surface of a comet thanks to Stardust-NEXT which on
14 February 2011 flew past 9P/Tempel 1 showing evident differences in some geological features
with respect to the first flyby of the Deep Impact spacecraft on 4 July 2005 (Veverka, 2012).
The Rosetta mission has provided the first ever escort of a comet along its orbit around the Sun,
allowing us to follow the orbital evolution of the surface during the perihelion passage.
The aim of the present work is the analysis of the water ice patches on the surface of 67P showing
the presence of coarser grains, which have been suggested as more stable in time (Filacchione et
al., 2016a; Barucci et al., 2016). In particular, here we focus on the spectral parameters, the
physical properties, and their evolution in time, using the hyperspectral data acquired by the
VIRTIS instrument on-board Rosetta.
The VIRTIS (Visible InfraRed and Thermal Imaging Spectrometer) instrument is described in detail
by Coradini et al. (2007); summarizing, VIRTIS is composed of a high spectral resolution channel
(VIRTIS-H) and a hyperspectral mapper (VIRTIS-M) covering the spectral range 0.22 - 1.04 μm
(VIRTIS-M-VIS) and 0.95 - 5.06 μm (VIRTIS-M-IR) with two detectors both with 432 spectral bands.
The present work takes advantage of the data acquired simultaneously by VIS and IR channels of
the VIRTIS-M instrument.
Dataset description
We have taken into account all the water ice patches discussed by Filacchione et al. (2016a) and by
Barucci et al. (2016). We run a search in the full VIRTIS observations database to retrieve all
observations covering the regions where icy patches were observed. The acquisitions with
unfavourable viewing geometry (incidence or emission angle > 70°), or with limited spatial
coverage were discarded. The patches with a limited temporal coverage were also discarded. In
addition, all pixels in shadow were excluded from the analysis.
Our final selection included three large patches: the "BAPs" (bright albedo patches) discussed by
Filacchione et al. (2015): BAP1 (longitude: 118°E, latitude: 13°N), BAP2 (longitude: 180°E, latitude:
-4°N), and the "SPOT 6" (longitude: 72°E, latitude: 3°N) discussed by Barucci et al. (2016). The
surface portions that are related to patches are covered by ~50 - ~5000 pixels respectively from
the least to the highest resolved acquisitions.
The measured spectra have been cleaned from artifacts and spikes (Raponi, 2014). The VIS and IR
channels have been co-registered to bridge the spectra of the two channels.
The VIRTIS-M dataset considered in this work is summarized in Tables 1, 2, and 3.
Spacecraft
clock time
Resolution
(m/px)
Local
solar
Average
incidence
angle (°)
Average
emission
angle (°)
Average
phase
angle (°)
Days from
2 Sept 2014
Helio
centric
(first acquisition)
distance (AU)
12.7
14.0
7.1
2.2
4.4
5.8
7.3
7.2
20.4
368245714
368291071
369364114
371998843
373202012
373248812
375524971
375613471
383518966
Table 1. Dataset selected for BAP 1: The first column indicates the spacecraft clock time which is also the
name of the acquisition. The diverse resolutions are due to the different distance of the spacecraft from the
comet surface at the time of the acquisitions. The column "local solar time" is defined according to the
longitude position, and it is in hour unit, being the comet rotation period divided in 24 hours.
0.000
0.525
12.944
43.439
57.365
57.906
84.251
85.275
176.77
3.443
3.440
3.365
3.174
3.084
3.080
2.904
2.897
2.230
56.4
60.7
58.4
58.3
67.0
61.1
63.7
67.6
53.0
time (h)
10.0025
9.3979
10.1517
9.3045
8.0281
8.9563
8.2379
7.6880
7.9924
48.1
29.3
44.9
55.0
40.5
42.0
44.7
53.2
45.0
38.8
45.2
67.5
95.0
96.2
94.7
97.8
97.8
51.7
Spacecraft
clock time
Resolution
(m/px)
Local
solar
12.5
12.5
15.0
7.7
5.0
19.2
367617963
367621623
368299474
369369214
377184571
384381667
Table 2. Same as Table 1 but for BAP 2.
time (h)
12.9488
14.4624
17.6636
17.0423
16.0833
19.7250
Spacecraft
clock time
Resolution
(m/px)
Local
solar
Average
incidence
angle (°)
Average
emission
angle (°)
Average
phase
angle (°)
Days from
25 Aug 2014
Helio
centric
(first acquisition)
distance (AU)
62.5
58.0
64.2
60.2
53.7
69.2
56.0
40.8
60.6
66.2
55.8
34.0
37.0
36.7
47.4
67.9
92.2
47.3
0.000
0.043
7.888
20.269
110.726
194.026
3.485
3.486
3.439
3.364
2.770
2.153
Average
incidence
angle (°)
Average
emission
angle (°)
Average
phase
angle (°)
Days from
24 Aug 2014
Helio
centric
(first acquisition)
distance (AU)
12.9
12.9
7.1
8.2
5.9
19.9
time (h)
13.9854
15.3450
11.9664
14.2728
14.9415
12.8080
26.5
33.8
31.5
56.9
49.3
52.5
39.1
38.8
68.9
108.8
91.0
51.9
50.7
55.3
57.4
55.9
53.3
31.9
3.488
3.488
3.364
3.059
2.842
2.233
0.000
0.043
20.66
68.785
100.846
184.04
367589075
367592759
369374074
373532092
376302211
383489908
Table 3. Same as Table 1 but for SPOT 6.
The two BAPs have been identified in correspondence of elevated structures, some of them
showing circular shapes, which display erosion and mass wasting on their sides. Both areas are
exposed towards the lower Imhotep (Thomas et al., 2015) plain where the waste material is
accumulated as boulders and debris (Filacchione et al., 2016a), which suggests that they originated
from mass wasting. SPOT 6 is a cluster of some twenty small bright spots located in the Khepry
region at the base of a scarp bordering a roundish flat terrace (Barucci et al., 2016). Also in this
case it is very probable that the origin of the icy region is due to mass wasting from the nearby flat
terrace.
The locations of the selected patches on the comet surface are indicated in Figure 1.
Figure 1. The three icy regions taken into account in this work are marked on the shape model of the comet
surface (Preusker et al., 2015). Imhotep and Khepry regions are indicated and colored on the shape model.
Analysis of spectral parameters
Spectral analysis of VIRTIS data has shown that spectra across the surface present a positive ("red")
VIS-IR slope and are basically featureless, except for the presence of a broad absorption feature
around 2.9-3.6 μm. This absorption band is ubiquitous on the entire comet surface, and it is
compatible with carbon-bearing compounds (Capaccioni et al., 2015). At large scale, most of the
surface is depleted of water ice except for some specific regions where the presence of water ice is
revealed by the shift of the absorption band toward shorter wavelengths (the band center moves
from 3.2 to 3.0 μm), a slight increase in the band depth, a change of slope (the spectra are flatter,
or "bluer") and a slight increase in albedo, which makes these spectra compatible with the
presence of few percent of fine-grained water ice (De Sanctis et al., 2015; Filacchione et al., 2016b;
Ciarniello et al., submitted).
By using both VIRTIS and OSIRIS data, Filacchione et al. (2016) and Barucci et al. (2016) observed a
number of wider and more complex icy patches. For these icy regions, all the diagnostic absorption
bands of water ice located at 1.05, 1.25, 1.5, 2.0, 3.0 μm (see Figure 2) are observed, and all the
spectra display bluer levels of slopes in the VIS and IR ranges from the ice-free areas.
Figure 2. Spectra of pure water ice simulated with Hapke model (see Appendix) by means of optical
constants provided by Warren et al. (1984), Mastrapa et al. (2008), Mastrapa et al. (2009), Clark et al.
(2012), for different grain sizes. The diagnostic absorption bands of water ice located at 1.05, 1.25, 1.5, 2.0,
3.0 μm are clearly visible.
In order to derive quantitative information on the amount of ice, we have identified a number of
spectral indicators that allow minimizing the spectral dimensionality, still preserving the relevant
information contained in the spectra. The spectral indicators are the slopes in VIS and IR, and band
areas.
The slope in the visible and in the infrared part of the spectra are calculated with a linear fit to the
spectrum in the range 0.55 – 0.75 μm for the visible channel, and 1.0 – 2.35 μm for the infrared
channel as shown in Figure 3.
The absorption band at 1.05 and 1.5 μm are not taken into account in this work because they are
prone to artifacts, being the former located close to the junction between the two spectral
channels, and the latter being affected by the junctions of the instrumental order sorting filters.
The band areas of the absorption bands at 1.25, 2.0 and 3.2 μm (see Figure 4) are calculated as:
𝑏
∫ 1 −
𝑎
reflectance(λ)
continuum(λ)
dλ
Being a and b the edges of the band: 1.16 - 1.38 μm, 1.83 - 2.24 μm, 2.62 - 3.60 μm, respectively.
The continuum used to characterize the bands is calculated with a linear fit between a and b. The
reflectance level correspondent to the band edges is calculated with a median of 7 spectels.
In those pixels where water ice is scarce or absent, the lowest value of the band area can be
negative because the shape of the spectrum is concave while the continuum defined by the linear
fit is a straight line. This is mainly the case for the band area at 1.25 μm. This does not affect our
spectral analysis.
In the region 2.9-3.6 μm there is a superposition of the ubiquitous organics band and the water ice
band. Although a variation of the band area at 3.0 μm can be due to either compound, we ruled
out the contribution of the organics, as the increase in band area was always associated with a
shift of the band center towards shorter wavelengths which clearly indicates an increasing content
of water ice rather than organic material as already observed by Filacchione et al. (2016b) and
Ciarniello et al. (submitted).
We mapped the spectral indicators across the selected areas, and in Figure 5 we show the
distribution of the three band areas and the spectral slopes for a single acquisition of BAP 1. The
spectral indicators are distributed according to different and well-defined patterns. The pattern
shown by the band area at 2.0 μm is intermediate between the pattern shown by the band areas
at 1.25 μm and 3.0 μm. The 1.25 μm band traces the areas with the largest ice abundance or
largest grain size, while the 3.0 μm band is more widespread. This will be described in more detail
in the next section.
The scatter plots in Figure 6 indicate the consistency of the results obtained for the various spectral
indicators; the increase in the 2.0 μm band area is associated to a flatter IR spectrum pointing to
an increase of the ice abundance, the same is true for the VIS and IR spectral slopes which
consistently decrease with increasing abundance of ice, as described in next section.
Figure 3. Reflectance spectrum of 67P normalized at 2.8 µm. This spectrum does not exhibit the typical
water ice features. The spectrum in the range 0.5-1.0 μm is measured by the VIS channel, while the 1-5 μm
range is covered by the IR channel. The broad absorption band at 3.2 μm is compatible with the presence of
organic compounds. The slopes are calculated with a linear fit to the spectrum in the range 0.55 – 0.75 μm
for the visible channel (blue line), and 1.0 – 2.35 μm for the infrared channel (red line).
Figure 4. A reflectance spectrum normalized at 1.1 μm, showing significant water ice absorption bands, is
depicted. The continuum in correspondence of the absorption bands at 1.25, 2.0 and 3.2 μm is calculated
with a linear fit at fixed wavelengths. The band area under the continuum is highlighted. The gray boxes
represent the spectral ranges affected by the junctions of the instrumental order sorting filters, which
produce unreliable spectral features, and thus are not taken into account in the analysis.
Figure 5. Analysis from acquisition 371998863 of BAP 1 (see Table 1). Panel A: context image (IR channel at
1.8 μm). Panels B, C, and D show the band areas (μm) of the absorption band at 1.25 μm, 2.0 μm and 3.0
μm, respectively. Panels E and F show the slopes (μm-1) respectively in the VIS and IR channel. Non-
illuminated pixels are marked in black and excluded from the analysis.
Figure 6. Scatter plot of the all pixels in acquisition 371998863, excluding those in shadow. Left and right
panel show the good correlation between the slope in the IR channel with respectively the band area at 2.0
μm and the slope in the VIS channel.
Spectral modeling
To quantify the ice abundance, grain size and mixing modalities we have applied Hapke's radiative
transfer model (Hapke, 2012) as described by Ciarniello et al. (2011). A similar approach was
adopted in the two previous papers by Filacchione et al. (2016) and De Sanctis et al. (2015).
The model allows deriving the spectrophotometric properties of mixtures of several end-members
characterized by their abundances and grain sizes. As the objective of this work is the study of
water ice, we minimized the number of end-members by using only crystalline water ice and a
"dark terrain" unit corresponding to the average spectrum of the comet's surface after the
application of photometric correction (Ciarniello et al., 2015). Crystalline ice has been simulated
using optical constants measured at T=160 K between 0.5 and 4 μm (Warren et al., 1984; Mastrapa
et al., 2008; Mastrapa et al., 2009; Clark et al., 2012).
In addition we adopted two mixing modalities: a linear (areal) mixing in which the surface is
modeled as a collection of patches of either pure water or pure dark terrain, and each photon
interacts only with a single species. An intimate mixing in which single grains of the end-members
are in contact with each other and the single photons can directly interact with both species.
The abundance and grain size of the water ice are free parameters. The abundance retrieved by
the model has to be intended as the fraction of total cross section of the ice grains over the area
subtended by the pixel.
Quantitative estimations on the deeper layers of ice cannot be directly retrieved because the
reflected light measured by the instrument interacts only with the first layer of ice grains.
Abundance and grain size were retrieved using an iterative best-fit procedure to minimize the
deviation between the model results and the measured spectra (see Appendix for further details
on the model and the best-fitting procedure).
The results, in agreement with our previous work (Filacchione et al., 2016) are that the observed
spectra require both areal and intimate mixture and, most interestingly, they require a bimodal
distribution of icy grains: mm-sized grains in areal mixtures, and ~50 μm sized grains in intimate
mixtures.
The left panel of Figure 7 shows an example of a fitted spectrum rich in water ice, which requires
0.9% water ice with 1800 μm grain size in areal mixture, and 2.2% water ice with 45 μm grain size
in intimate mixture with the dark material. To illustrate the effect of the mixing modalities on the
right panel we show the spectra produced by the two different populations. The most striking
differences are the absence of the 1.25 μm band in the small grain size mixture (blue curve) and
the saturation of all ice bands in the large grain spectrum (red curve).
In order to characterize the influence of ice abundance and grain sizes of both grain populations on
the overall shape of the spectra we performed specific simulations by taking into account the
typical retrieved values, as shown in Figure 8. The simulated spectra are normalized in order to
highlight the variations of the spectral indicators as a function of the parameters of the model:
- The slope in the VIS channel is only affected by variation in abundance for both populations of
grains;
- The slope in the IR channel is affected by all four parameters;
- The band area at 1.25 μm is only affected by properties of the ice with large grains in areal
mixture;
- The band area at 3.0 μm is only affected by properties of the ice with small grains in intimate
mixture;
- Band area at 1.5 μm (not taken into account in this work) and band area at 2.0 μm are affected by
all parameters.
While it's relatively common to find ice patch spectra without the 1.25 μm absorption band (only
intimately mixed water ice), spectra which clearly present a saturation of the icy features (see red
curve of Figure 7) are very uncommon or absent among the pixels constituting the analyzed areas.
This means that the patches of mm-sized water ice are usually embedded in a surface of intimately
mixed and finer grained ice. This is indeed what is observed in Figure 5 (panels B and D).
In Figure 9 two different collections of measured spectra show the effect of a variation of the
abundance of both populations of water ice. Model used to retrieve the abundances is the same
used to produce the simulated spectra in Figure 8.
The variety of the measured spectra can be mainly attributed to a variation of the abundance and
partially to the grain size of the water ice within the spots. This is shown in the scatter plots in
Figure 10, which reports the case of the acquisition 371998863 (the same as for Figures 5 and 6) as
an example: the abundance of water ice in areal mixture (panel A of Figure 10) is in good
correlation with the band area at 1.25 μm. By grouping the symbols in the Figure according to their
grain size it seems that for the pixels with grain size < 2000 μm the upper limit is ~1% of
abundance, while it is higher for pixels with larger grain size. Panels B, C, D of Figure 10 show the
correlation between the abundance of the water ice in intimate mixture with the band area at 2.0
μm, the band area at 3.0 μm, and the slope in the IR channel, respectively. For the 2.0 μm band
area and the IR slope case, the different grain sizes lead to different lines of correlation. For the 3.0
μm band area, the different groups seem to follow the same line, but they are limited by different
upper bounds. These three parameters are marginally affected by the variation of properties of the
mm-sized grain population because of the lower amount of the latter. For the purpose of this
analysis only spectra with a band area at 2.0 μm larger than 0.005 μm are taken into account. This
threshold is equal to the mean level + 3 standard deviations of the band area, as calculated in a
region outside of the icy patches. Possible interpretations of the results obtained here are
discussed in the next sections under the general context of the temporal evolution of the icy
patches.
Figure 7.Left panel: VIRTIS reflectance spectrum (sample 25, line 20, spacecraft clock time 371998843) in
black. Error bars are calculated as described by Raponi (2014). Best fit in red: 0.9% water ice with 1800 μm
grain size in areal mixture, and 2.2% water ice with 45 μm grain size in intimate mixture with the dark
material. Spectral ranges affected by the junctions of the filters are not taken into account in the analysis.
Discrepancy between best fit and the VIS channel can be due to the not perfect co-registration between VIS
and IR channels. In the right panel the two populations of the water ice retrieved with the model have been
decoupled: in red a scaled spectrum of pure water ice with 1800 μm grain size, in blue an intimate mixture
of 2.2% water ice with 45 μm grain size and 97.8% dark terrain.
Figure 8. Simulated reflectance spectra are shown to highlight the effect of variation in abundance and grain
size of the water ice. The black curve represents an intimate mixture of dark terrain and 2% of water ice
with 50 μm grain size, which in turn is in areal mixing with 0.5% water ice with 2000 μm grain size. The red
and blue curves are simulated by varying one of these parameters at a time, as indicated: the panels on the
left show the effect of a variation in abundance, and the panels on the right a variation in grain size, for
both populations of water ice: intimate mixing in the upper panels (spectra normalized at 1.16 μm), and
areal mixing in the lower panels (spectra normalized at 2.62 μm). The spectral ranges are not taken into
account in correspondence of artifacts that have prevented the definition of the albedo of the dark terrain.
Figure 9. According to the model, the collected measured spectra show the effect of a variation of the
abundance of water ice in intimate mixture (left), and the effect of variation of the abundance for the areal
mixtures (right). The retrieved abundance is shown in the panels.
Figure 10. Analysis of acquisition 371998843. Spectra selected are those with band area at 2.0 μm larger
than 0.005. Panel A: Abundance of water ice in areal mixture as a function of the band area at 1.25 μm.
Panels B, C, D show the correlation between the abundance of water ice in intimate mixture with
respectively the band area at 2.0 μm, the band area at 3.0 μm, and the slope in the IR channel. The symbols
are marked with different colors according to their grain size as indicated in the panels.
Water ice property retrieval by means of spectra in the visible range
The aim of this section is to give a useful guideline for the retrieval of the icy patches' properties by
means of instruments operating only in the visible spectral range.
Thanks to the Hapke model we can retrieve the water ice abundance by modeling the normalized
spectral slope. We do not take into account the absolute signal level because it can be affected by
uncertainties on the radiometric and photometric accuracy as well as errors on the local geometry
information, due to unresolved shadows and roughness.
As shown in Figure 8 the slope in the VIS channel is sensitive to the water ice abundance and the
type of mixing (areal or intimate), while it is not affected by the grain size. In principle we cannot
distinguish between the two types of mixing by just the slope in the VIS channel, and thus we can
only give lower and upper limits by assuming water ice in areal or intimate mixture with the dark
material, respectively (black and red lines in Figure 11). However, from the analysis performed on
the three icy patches discussed in the present work, we obtain as a general result that the ratio
between the abundances of water ice in intimate and areal mixture is in the range 1 - 10 (see next
sections and upper panels of Figures 15, 16, and 17). By assuming this range of ratios, we can
reduce the gap between the lower and upper limit (blue and green line in Figure 11).
The phase function should also be accounted for by the model, as it can produce slope variations
of the order of +0.007 μm-1 / deg, according to Ciarniello et al. (2015).
Moreover, the dark terrain used as a spectral end-member could contain itself a small percentage
of water ice, and thus the water ice abundance could be underestimated.
Figure 11. Curves: theoretical abundance of water ice as a function of the slope in the VIS channel, as
calculated by means of the Hapke radiative transfer model at a phase of 95.0° (same phase as acquisition
371998843). The red and black lines correspond to water ice entirely in intimate or in areal mixture with the
dark terrain, respectively. The blue and red lines correspond to the ratio between the amount of water ice
in intimate and areal mixture equal to 1 and 10, respectively. Dots: Water ice total abundance as a function
of the measured slope in the VIS channel, for the acquisition 371998843. Spectra selected are those with
band area at 2.0 μm larger than 0.005. Abundance is retrieved as described in previous section by means of
the spectra in the IR channel. Dashed gray line: average slopes of the dots. The width of the dashed line is
equal to the error on the average slope.
We take into account the acquisition 371998843 in order to compare the theoretical curves shown
in Figure 11 with the abundance retrieved as in previous section by means of the spectra in the IR
channel (the points of Figure 11). The ratio between the water ice abundance in intimate and areal
mixing is close to 10 for this acquisition (see upper panels of Figure 15). This is consistent with the
dispersion of the points shown in Figure 11 around a curve close to the green one.
By taking the average of the VIS slope obtained for the points shown in Figure 11 we can constrain
the amount of water ice of the patch in the range 0.6 – 1.2 %, assuming the ratio between the
abundances of water ice in intimate and areal mixture in the range 1 - 10. The total abundance
retrieved with spectra in IR range is 1.1 % (see upper left panel of Figure 15), thus demonstrating
the consistency of the result obtained with spectra in the VIS channel with that obtained with
spectra in the IR channel.
Temporal evolution of ice-rich areas
The collected acquisitions listed in Tables 1, 2, and 3 allow following the temporal evolution of the
three icy regions analyzed in this work. According to the previous discussion and as shown in Fig. 4,
we selected the absorption band at 2.0 μm as a reliable indicator of the water ice presence. Each
hyperspectral cube was orthorectified and projected on a cylindrical map. The temporal evolution
of the 2.0 μm band area in each ice-rich region, BAP 1, BAP 2 and SPOT 6, is shown in Figures 12,
13 and 14.
To give a quantitative estimate of the amount of ice present in the region under observation, the
spectra with band area at 2.0 μm larger than 0.005 are taken into account. The retrieved water ice
abundances are averaged, weighted by the area covered by each pixel, calculated as Area =
resolution2 / cos(emission angle). The result is the fraction of the total area of the patch in which
the solar flux interacts with the icy grains, which is a proxy for the surface density of water ice
(upper left panels of Figures 15, 16 and 17). This abundance is multiplied by the total area
subtended by the pixels taken into account, to obtain the total cross section of the water ice
(upper right panels of Figures 15, 16 and 17). The lower panels of Figures 15, 16, and 17 show the
grain size retrieved for the two populations of grains as a function of time.
Figure 12. The band area at 2.0 μm is mapped for each acquisition of BAP 1. The panels represent the
acquisition listed in Table 1. The increase of the extension and intensity of the feature in the first 6 panels,
and then their decrease are quite evident. In the last panel the feature is very weak or absent. The dark
areas are in shadow. The white areas outside the edge of the map in panels D and F are outside the field of
view of the acquisitions.
Figure 13. Same as Figure 12 but for BAP 2.
Figure 14. Same as Figure 12 but for SPOT 6.
Figure 15. BAP 1: water ice abundance (upper left panel) and total cross section of water ice (upper right
panel) is plotted as a function of time for the two types of mixtures modeled. The latter could be affected
by the limited view of the icy region because of shadows, or because part of the region is outside the field
of view of the instrument: those symbols are marked with an arrow indicating a lower limit. The lines are
produced with a polynomial fit of the points (without including those with the arrow). The bell shaped
curve is evident for both populations of water ice. The total amount of the water ice in areal mixture is
always a small portion of the total amount of water ice which covers the surface. The errors on the
retrieved values (not shown for the sake of clarity) are of the order of the size of the symbols.
Lower panel: the mean grain size is plotted as a function of time.
The errors are calculated as the standard deviation of the values divided by the square root of the number
of selected pixels.
Figure 16. Same as Figure 15 but for BAP 2.
Figure 17. Same as Figure 15 but for SPOT 6.
Discussion
The most striking aspect of the observations presented in this work is that all three ice-rich regions
have a life cycle in which the icy patch increases in size and abundance, reaches a maximum and is
followed by a rapid decrease after which, eventually, the ice disappears from the surface. The
three areas analyzed are all very close to the equator, which is consistent with their similar thermal
history and evolution.
The determined lifecycle allows us to realistically exclude a diurnal cycle as the major driver for the
observed variability and, indeed as reported in Tables 1, 2 and 3, the observations cover a
restricted range of local times and incidence angles.
However, significant variations on smaller time scales can be appreciated both in the extension and
intensity of the icy features. As an example panels G and H of Figure 12 concern two acquisitions of
BAP 1 which are close in time (24 hours, two comet's rotations) but the latter presents a
significantly smaller extension and lower intensity of the 2.0 μm band. This is also confirmed by
the retrieved abundances (upper panels of Figure 15).
This difference cannot be accounted for entirely by the diverse shadowing. Although we cannot
exclude uncertainties due to the different resolution and viewing geometry of the acquisitions, it is
reasonable to explain such variation with a diurnal cycle, which produces short-term behavior,
overlapped to the seasonal cycle. The spatial distribution of the 3.0 μm and 1.25 μm band area
values could support the understanding of the temporal stability of the patch: as discussed above,
the 3.0 μm band area (panel D of Figure 5) is an indicator of the presence of the finer intimately
mixed icy grains. Their concentration seems to follow the shadow, being larger closer to the non-
illuminated areas, indicating a sublimation of the finer icy grains when exposed to the solar light.
Conversely, the 1.25 μm band area is an indicator of larger grains in areal mixing (panel B of Figure
5), which are grouped in small spots independently from the shadow position, indicating their
temporal stability with respect to the solar input, at least in the diurnal time scale.
Most of the water ice is in the form of finer grains, in intimate mixture with the dark material. This
population is the main driver of the variability of the amount of water ice (see upper panels of
Figures 15, 16, 17). In contrast, ice in mm-sized grains appears very locally concentrated in specific
spots embedded in the more widespread finer grains.
Except for BAP 1, the average size of the mm-sized grain population seems to decrease as a
function of time (lower left panels of Figures 15, 16, 17), most probably because the sublimation
produces erosion of the grains. Conversely, the average size of the ~50 μm sized grain population
increases with time for all patches. This can be explained by the complete sublimation of the
smaller grains, which causes the size distribution to shift towards larger sizes. This is also
confirmed by their spatial distribution shown in panel C of Figure 10: the pixels with the larger
amount of water ice (blue and black dots) are populated by the finer grains and vice versa,
indicating that where the sublimation has been less effective the water ice is still abundant with a
small average grain size. On the other hand, inside the icy spots with mm-sized grains, more
affected by grain erosion than complete sublimation, the pixels present larger abundances and
larger grain sizes (green dots of panel A in Figure 10) where the solar input has been less intense.
Conclusion
As reported by Filacchione et al. (2015) and Barucci et al. (2016) the observed patches occur in
areas affected by mass wasting and erosion; consequently, the three ice-rich regions very likely
represent the occasional exposure of sub-surface layers fairly rich in water ice. This is a common
finding on 67P where ice is predominantly absent and only temporary present at the surface, when
exposed due to occasional erosional events or in form of recondensed ice in specific location (De
Sanctis et al, 2015).
In addition, in all these areas, along with ice grains of tens of microns intimately mixed with the
dark component, we observed also areas with mm-sized grains, which make a relevant
contribution to the overall water ice abundance. The presence of mm-sized grains is quite
unexpected at the surface of the nucleus because, as reported by many authors (Yang et al., 2009;
Sunshine et al., 2007; Protopapa et al., 2014; De Sanctis et al., 2015; Ciarniello et al., submitted),
modeling of observed coma and nucleus surface spectra seems to indicate that the typical size
range of the ice grains observed in comets never exceeds tens of micron.
Filacchione et al. (2015) already identified mm-sized ice grains in BAP 1 and BAP 2. The addition of
SPOT 6 from a completely different region indicates that the presence of larger grains in subsurface
layers is a common feature of the water ice's physical status. The large size possibly derives from
physical processes acting within deeper layers of the nucleus, as for instance by the growth of
secondary ice crystals from vapor diffusion in ice-rich colder layers, and/or by ice grain sintering
(see Filacchione et al., 2016a, and references therein).
The exposure of deeper layers is consistent with their occurrence in "active" areas where falls or
landslides could have caused the occasional exposure of water ice rich layers.
After the initial exposure of the ice, the activity of the affected area increases thus causing dust
removal powered by sublimation, which provides a positive feedback on the exposure itself. The
process develops as the solar flux increases, and it reaches a turning point when the exposure rate
is overcome by the sublimation rate, until the complete sublimation of the patch.
The similarity in the lifecycles of the three patches analyzed in this work (180 days) and the
similarity in the ice grains size distribution, seems to indicate that water ice is rather evenly
distributed in the subsurface layers and mixed to the ubiquitous dark terrain. Furthermore, our
present findings also indicate the absence of large water ice reservoirs in the subsurface. This is
confirmed by the recent work of Filacchione et al, 2016b and Ciarniello et al., 2016 where an
overall and diffuse increase in the abundance in water ice has been derived from the VIRTIS
observations spanning the period August 2014 through May 2015. Indeed, the consistent erosion
rate of the nucleus of 67P (Bertaux, 2015; Keller et al., 2015), would allow exposure of large-scale
dis-homogeneities, and in particular of regions highly enriched in water ice, which would last for a
longer timescale, if present.
The knowledge of the temporal evolution of the ice-rich regions identified, can constrain the
thickness of the layer and the amount of water released to the coma during the sublimation
process: this will be the subject of a following work.
Appendix
Hapke model and fitting procedure description
A quantitative spectral analysis of the composition has been performed using Hapke's radiative
transfer model (Hapke, 2012) as described by Ciarniello et al. (2011).
Eq. 1A 𝑟(𝑖, 𝑒, 𝑔) =
𝑤(𝜆)
4𝜋
µ0
µ0 + µ
[𝐵𝑆𝐻 𝑝(𝑔, 𝜆) + 𝐻(𝑤, µ0)𝐻(𝑤, µ) − 1] 𝑆(𝑖, 𝑒, 𝑔, 𝜃)𝐵𝐶𝐵
Where r is the bidirectional reflectance, i, e, g are the incidence, emission and phase angle,
respectively; w is the single scattering albedo (SSA); p(g,λ) is the single particle phase function; μ0,
μ are the cosines of the incidence and emission angles; H(w,μ) the Ambartsumian-Chandrasekhar
functions describing the multiple scattering components; BSH the shadow hiding opposition effect
term; BCB the coherent back-scattering opposition effect; S(i, e, g, θ) the shadow function modeling
large scale roughness and θ the average surface slope.
During the VIRTIS-M observations described in this paper the solar phase was always > 35°. As a
consequence, the shadow hiding effect is negligible. This implies posing the terms BSH and BCB
equal to 1. The other photometric parameters in the equation are fixed as resulting from Ciarniello
et al. (2015).
The geometry information is derived from the information on the digital shape model of the comet
(Preusker et al., 2015), the spacecraft attitude, the comet's attitude, and the relative positions.
The single scattering albedo w can be modelled as a mixing between the dark terrain of the comet
and water ice. The former was defined by Ciarniello et al. (2015), for the latter the SSA is calculated
from the optical constants (Warren et al., 1984; Mastrapa et al., 2008; Mastrapa et al., 2009; Clark
et al., 2012) and from the grain size by means of the Hapke compositional model (2012).
We have used areal and intimate mixing modalities in the simulations: in the areal mixing the
surface is modelled as patches of pure water ice and dark terrain, with each photon scattered
within one patch. In this case the resulting reflectance is a linear combination of the reflectance of
the different end-members.
In the intimate mixing model the particles of the two end-member materials are in contact with
each other and both are involved in the scattering of a single photon. In intimate mixing the single
scattering albedo of the mixture is the weighted average, through their relative abundance P, of
the Dark Terrain (DT) and Water Ice (WI) single scattering albedos.
Most of the spectra inside the patches require both kinds of mixtures and two different grain sizes
of water ice. Thus the resulting reflectance r(λ) is given by:
Eq. 2A r(λ) = rWI_A(λ) PWI_A + rWI_DT_I(λ) (PWI_I + PDT),
with: rWI_DT_I = rWI_DT_I (w(λ)DT PDT + w(λ)WI_I PWI_I),
Where r is the bidirectional reflectance (Eq. 1A) and the suffix "A" and "I" state for the areal and
intimate mixture.
The best-fitting result is obtained by applying the Levenberg-Marquardt method for nonlinear least
squares multiple regression (Marquardt, 1963)
Free parameters of the model are:
- Abundance and grain size of the water ice for both intimate and areal mixtures.
- A multiplicative constant of the absolute level of reflectance of the model, in order to account for
uncertainties on the radiometric and photometric accuracy as well as errors on the local geometry
information, due to unresolved shadows and roughness.
- A slope added to the model in order to better fit the measured spectrum in the IR channel: in
some cases, the measured spectra present an artificial slope where high signal contrast is
measured between adjacent pixels, like regions near shadows. This is due to the increasing full
with half maximum of the spatial point spread function toward longer wavelengths (Filacchione,
2006).
- Temperature and effective emissivity (Davidsson et al., 2009). The latter is the product of the
directional emissivity (see next section) and a free parameter used to account for unresolved
shadow and the structure of the surface (Davidsson et al., 2009). Its interpretation is outside the
scope of the present work.
The total radiance is modeled by accounting for both the contributions of the reflected sunlight,
and the thermal emission:
Eq. 3A Rad(λ) = r(λ)* F⨀ / D2 + εeff * B(λ, T)
ελ = H(w, μ) * γ(w)
Where r(λ) is the Hapke bidirectional reflectance (Eq. 1A), F⨀ is the solar irradiance at 1 AU, D is
the heliocentric distance (in AU), εeff is the effective emissivity, B(λ, T) is the Planck function.
Emissivity
In order to model the thermal emission we take into account the directional emissivity (ελ) relation
(Hapke, 2012) which is a function of the single scattering albedo (w):
Eq 4A
being μ the emission angle cosine, and H the Ambartsumian-Chandrasekhar function. According to
Kirchhoff's law of thermal radiation, the emissivity and reflectance spectral contrasts (difference
between continuum and band signal) have opposite signs. The higher is the thermal emission in
the spectral range of the absorption band, the lower is the band depth (by summing the
reflectance and the thermal emission), and it can even become an emission feature. By means of
Eqs. 1A-4A, the spectral contrast has been simulated for the sum of the reflected and thermally
emitted fluxes. In Figure 1A two different conditions are simulated on the base of the viewing
geometry of the icy patches analyzed in this work and the temperatures obtained by Filacchione et
al. (2015) and Barucci et al. (2016). As a result the emissivity spectral contrast can be considered
negligible with respect to the reflectance spectral contrast.
Figure 1A. The plots show the sum of the simulated spectral contrasts of reflected and emitted light. It is
normalized to the value at shorter wavelengths (< 3.0 μm) for which thermal emission is not observed. Two
conservative thermal conditions are simulated. For T = 180 K (blue line) the effect of the thermal emission is
negligible in the spectral range taken into account in this work. For T = 210 K (red line) a possible absorption
band at 5.0 μm would be flattened, and longward of that it would become an emission feature. Shortward
of 4.0 μm, which is the limit of our analysis, the spectral signatures are affected as much as 5%. For the
absorption band measured at 3.0 - 3.5 μm the effect is always negligible.
Uniqueness of the solutions
In order to produce the best fit we should evaluate if the solutions steadily converge, or if there
are secondary solutions. This evaluation is made by the investigation of the χ2 function defined in
the N-dimensional space, where N is the number of the free parameters. As an example we report
in Figure 2A the projections of this space around the best fit of Figure 5: the icy spectrum for which
the resulting water ice in areal mixture is 0.9 % and 1800 μm grain size, and 2.2 % at 45 μm grain
size of water ice in intimate mixture with the dark terrain. As a result of these investigations
significant secondary minima have never been found. Thus we can rely on the uniqueness of the
solution.
Figure 2A. In the four panels the χ2 function is plotted as a function of water ice abundances and grain sizes
of the two populations of water ice. The absence of secondary minima confirms that the solutions are
stable and there are no other sets of parameters which can produce models as good as the one retrieved.
Acknowledgements
We thank the following institutions and agencies for support of this work: Italian Space Agency
(ASI, Italy) contract number I/024/12/1, Centre National d'Etudes Spatiales (CNES, France), DLR
(Germany), NASA (USA) Rosetta Program, and Science and Technology Facilities Council (UK).
VIRTIS was built by a consortium, which includes Italy, France, and Germany, under the scientific
responsibility of the Istituto di Astrofisica e Planetologia Spaziali of INAF, Italy, which also guides
the scientific operations. The VIRTIS instrument development, led by the prime contractor
Leonardo-Finmeccanica (Florence, Italy), has been funded and managed by ASI, with contributions
from Observatoire de Meudon financed by CNES, and from DLR. We thank the Rosetta Science
Ground Segment and the Rosetta Mission Operations Centre for their support throughout all the
phases of the mission. The VIRTIS calibrated data will be available through the ESA's Planetary
Science Archive Website (www.rssd.esa.int/index.php?project=PSA&page=index) and is available
upon request until posted to the archive.
References
A'Hearn M. F. et al., 2005, Science, 310, 258-264.
A'Hearn M. F. et al., 2011a, Science, 332, 1396-1400.
A'Hearn M. F. et al., 2011b, Annu. Rev. Astron Astroph, 49, 281-299.
Barucci M. A. et al., 2016, submitted to A&A.
Bertaux J.-L., 2015, Astronomy & Astrophysics, 583, id.A38, 10 pp.
Bockelée-Morvan D. and Rickman H., 1997, Earth Moon Planets 79, 55-57.
Bockelée-Morvan D. et al., 2004, the university of Arizona press, Tucson, 391-423.
Capaccioni F. et al., 2015, Science 347, aaa0628.
Ciarniello M. et al., 2011, Icarus 214, 541-555.
Ciarniello M. et al., 2015, A&A, 583, 31.
Ciarniello M. et al., 2016, MNRAS, submitted to this special issue.
Clark R. N. et al., 2012, Icarus, 218, 831.
Coradini A. et al., 2007, Space Sci. Rev., 128, 529.
Davidsson B. J. R. et al, 2009, Icarus, 201, 335-357.
De Sanctis M. C. et al, 2015, Nature, 525, 500.
Feaga L. M. et al., 2007, Icarus, 190, 345-356.
Filacchione G., 2006, PhD Thesis, Università di Napoli Federico II
Filacchione G. et al., 2016a, Nature, 529, 368
Filacchione G. et al., 2016b, Icarus, 274, p. 334-349.
Hapke B., 2012, Theory of reflectance and emittance spectroscopy. Cambridge University Press.
Keller H. U. et al., 2007, Space Sci. Rev., 128, 26
Keller H. U. et al., Astronomy & Astrophysics, 583, id.A34, 16 pp.
Marquardt. J., 1963,Soc. Ind. Appl. Math. 11(2), 431–441.
Mastrapa R. et al., 2008, Icarus, 197, 307-320.
Mastrapa R. et al., 2009, Astrophys. J., 701, 1347-1356.
Mumma M. J. and Charnley S. B., 2011, A&A 49, 471-524.
Pommerol A. et al. 2015, A&A, 583, 25.
Preusker F. et al., 2015, A&A, 583, 19.
Protopapa S. et al., 2014, Icarus, 238, p. 191-204.
Raponi A. et al., 2013, LPI Contribution No. 1719, p.1507.
Raponi A., 2014, PhD Thesis, Università degli studi di Roma Tor Vergata
Sunshine J. M. et al., 2006, Science, 311, 1453-1455.
Sunshine J. M. et al., 2007, Icarus 190, 284-294.
Thomas P. et al., 2015, Science 347
Veverka J. et al., 2012, Icarus, 222, 424-435.
Warren S. G. et al., 1984, Appl. Opt. 23, 1206.
Yang B. et al., 2009, Astron. J., 137, 4538-4546.
|
0907.4318 | 3 | 0907 | 2010-08-24T15:12:51 | Efficient computation of the quadrupole light deflection | [
"astro-ph.EP",
"astro-ph.IM"
] | Efficient computation of the quadrupole light deflection for quasars/quasars and solar system objects within the framework of the baseline Gaia relativity model (GREM) is discussed. Two refinements have been achieved with the goal to improve the performance of the model: First, the quadrupole deflection formulas for both cases are simplified as much as possible considering the Gaia nominal orbit (only approximate minimal distances between Gaia and the giant planets were used here), physical parameters of the giant planets and the envisaged accuracy of 1 microarcsecond for individual systematic effects. The recommended formulas are given by Eq.(40) for stars/quasars and by Eq.(81) for solar system objects. Second, simple expressions for the upper estimate of the quadrupole light deflection have been found allowing, with a few additional arithmetical operations, to judge a priori if the quadrupole light deflection should be computed or not for a given source and for a given requested accuracy. The recommended criteria are given by Eq.(45) for stars/quasars and by Eq.(92) for solar system objects. Additionally, the quadrupole Shapiro effect for solar system objects is reconsidered. A strict upper bound of quadrupole Shapiro effect for solar system objects is given in Eq.(109). | astro-ph.EP | astro-ph | Efficient computation of the quadrupole light deflection
Sven Zschocke, Sergei A. Klioner
Lohrmann Observatory,
Dresden Technical University,
Mommsen Str.
13,
D-01062 Dresden, Germany
GAIA-CA-TN-LO-SZ-001-2
October 31, 2018
Efficient computation of the quadrupole light deflection for both stars/quasars
and solar system objects within the framework of the baseline Gaia relativity model
(GREM) is discussed. Two refinements have been achieved with the goal to improve
the performance of the model:
-- The quadrupole deflection formulas for both cases are simplified as much as possible
considering the Gaia nominal orbit (only approximate minimal distances between
Gaia and the giant planets were used here), physical parameters of the giant planets
and the envisaged accuracy of 1 µas for individual systematic effects. The recom-
mended formulas are given by Eq. (40) for stars/quasars and by Eq. (81) for solar
system objects.
-- Simple expressions for the upper estimate of the quadrupole light deflection have
been found allowing, with a few additional arithmetical operations, to judge a
priori if the quadrupole light deflection should be computed or not for a given
source and for a given requested accuracy. The recommended criteria are given by
Eq. (45) for stars/quasars and by Eq. (92) for solar system objects.
The quadrupole Shapiro effect for solar system objects is reconsidered. A strict up-
per bound of quadrupole Shapiro effect for solar system objects is given in Eq. (109).
0
1
0
2
g
u
A
4
2
.
]
P
E
h
p
-
o
r
t
s
a
[
3
v
8
1
3
4
.
7
0
9
0
:
v
i
X
r
a
Contents
I. Introduction
II. Some basics about light propagation
III. The quadrupole light deflection for stars and quasars
IV. Approximation of the quadrupole light deflection formula for stars and
quasars
A. Estimate of the vectorial coefficients
B. Estimate of the scalar functions
C. Collection of all terms
V. An upper estimate of the quadrupole light deflection for stars and
quasars
VI. The quadrupole light deflection for solar system objects
VII. Approximation of quadrupole light deflection for solar system objects
A. Estimate of the vectorial coefficients
B. Estimate of the scalar funtions
1. Estimate of BA
2. Estimate of CA
3. Estimate of DA
C. Collection of all terms
VIII. An upper estimate of the quadrupole light deflection for solar system
objects
IX. Shapiro effect for solar system objects
X. Numerical tests
XI. Summary
References
A. Proof of Eq. (87)
B. Proof of Eq. (94)
C. Proof of inequalities (107) and (108)
2
3
3
5
6
7
8
8
9
10
11
11
12
12
13
13
14
14
16
17
19
19
21
21
22
3
I.
INTRODUCTION
Gaia mission will reach an accuracy on microarcsecond (µas) level. This level of accuracy
requires a precise modelling of light propagation.
In particular, the light deflection due
to quadrupole gravitational field of deflecting bodies should be taken into account [12].
Analytical formulas for quadrupole light deflection are well known. Analytical solutions of
light deflection in a quadrupole gravitational field have been investigated by many authors
[1, 6 -- 13, 15]. For the first time the full analytical solution for the light trajectory in a
quadrupole field has been obtained in [10]. These results were confirmed by a different
approach in [2]. Various generalization (higher-order multipole moments, time-dependence,
etc.) were derived in [3 -- 5]. The formulas suitable for high-accuracy data reduction are given
e.g. in [12].
However, the Gaia mission will determine the astrometric positions of about 109 objects,
implying a data reduction of about 1012 individual observations during the mission time
of 5 years.
It is, therefore, obvious that efficient analytical solutions of quadrupole light
deflection are mandatory in order to succeed with data reduction. But the full expressions
of these solutions of light deflection are rather involved and much too time-consuming for
practical Gaia data reduction. Furthermore, the quadrupole light deflection will reach the
microarcsecond level only for objects within a small observational field of giant planets.
Accordingly, it is highly useful to find analytical criteria by means of which one can decide
whether or not the quadrupole field needs to be taken into account. Such criteria can only
be obtained by simpler analytical formulas. And finally, the implementation of the full
expressions involve round-off errors in the Gaia data reduction. Therefore, it is tempting
to obtain simpler analytical expressions which will not be hampered by such problems. By
means of the assumption the Gaia spacecraft is located near the Earth's orbit (Gaia will have
a Lissajous-like orbit around Lagrange point L2), we have obtained simpler expressions valid
on microarcsecond level of accuracy. Criteria by means of which one can decide whether or
not it is necessary to take into account the quadrupole effect have been derived.
Furthermore, the accuracy of radio and laser radar links of future missions like Bepi-
Columbo or Juno require modelling of the light travel time at the level of millimeters.
Hence, also the Shapiro delay due to quadrupole fields is of practical interest. Therefore and
for reasons of completeness, the known analytical expressions of quadrupole Shapiro effect
will be reconsidered. Especially, an improved estimate of this effect is given.
The report is organized as follows: In Section II we summarize some basics about light
deflection and introduce the notation. In Section III the full quadrupole formula in post-
Newtonian order for stars and quasars is presented. A simplified expression and criteria for
stars and quasars for quadrupole light deflection are given in Sections IV and V. In Section
VI the full quadrupole formula in post-Newtonian order for solar system objects is presented.
A simplified expression and criteria for quadrupole light deflection of solar system objects
are given in Sections VII and VIII. A strict upper bound of quadrupole Shapiro effect is
given in Section IX. Numerical tests are given in Section X. The findings are summarized
in Section XI.
II. SOME BASICS ABOUT LIGHT PROPAGATION
Let us summarize some basic formulas of light propagation in post-Newtonian approxi-
mation. The geodetic equations in post-Newtonian order is linear with respect to the metric
4
components and, therefore, the coordinates of a photon and the derivative with respect to
coordinate time t is given by [12]
x(t) = x(t0) + c σ (t − t0) +Xi
x(t) = c σ +Xi
∆ xi(t) .
∆xi(t) ,
(1)
(2)
Here, t0 is the time moment of emission, x0 = x(t0) is the position of the photon at the
is the unit tangent
moment of emission, i.e. the position of source, and σ = limt→−∞
vector at infinitly past. The position of observer is x1 = x(t1) and t1 is the moment
of observation. The unit coordinate direction of the light propagation at the moment of
c
x(t)
observation reads n =
x(t1)
x(t1)
. In post-Newtonian order the transformation σ to n reads
where
n = σ +Xi
δσi + O(cid:16)c−4(cid:17) ,
δσi = σ ×(cid:16)c−1 ∆ xi (t1) × σ(cid:17) .
(3)
(4)
The sum in (3) runs over individual terms in the metric of various physical origins (e.g.
monopole gravitational field of various bodies, quadrupole fields, higher multipole fields,
etc.). The spherical symmetric term of light deflection is given by
δσpN(t1) = σ ×(cid:16)c−1 ∆pN x(t1) × σ(cid:17) = XA
σ · rA
1 ! .
pN(t1) =−
rA
A 1 +
(1 + γ) G MA
dA
d2
δσA
c2
1
δσA
pN(t1) ,
The sum in (5) runs over the bodies A of solar system. The impact vector
dA = σ ×(cid:16)rA
1 × σ(cid:17) ,
(5)
(6)
has been introduced having the absolute value dA = dA. The vector rA
1 = x1 − xA is
directed from body A towards the observer; the absolute value rA
1 , and γ is the PPN
parameter (for general theory of relativity γ = 1). We also note the absolute value of the
monopole contribution from one body A:
1 = rA
= (1 + γ)
G MA
c2
1
dA 1 +
1
σ · rA
1 ! ≤ 2 (1 + γ)
rA
G MA
c2
1
dA
.
(7)
(cid:12)(cid:12)(cid:12)
δσA
pN(t1)(cid:12)(cid:12)(cid:12)
In order to consider light propagation between two given points x0 and x1 (as it is needed
for the data processing for solar system objects) let us define the vector rA
0 = x0 − xA,
which is directed from body A towards the source and the vector R = x1 − x0 = rA
1 − rA
0
is directed from source towards observer. Furthermore, the unit direction from source to
observer is k = R
R , and the absolute value R = R, rA
0 = rA
0 .
In post-Newtonian order, the transformation k to n reads
where
n = k +Xi
δki + O(cid:16)c−4(cid:17) ,
5
(8)
(9)
δki = k ×(cid:16)c−1 ∆ xi (t1) − R−1 ∆xi (t1)(cid:17) × k .
The sum in (8) runs over individual terms in the metric of various physical origins (e.g.
monopole gravitational field of various bodies, quadrupole fields, higher multipole fields,
etc.). Thus, the spherical symmetric part reads (cf. Eq. (70) in [12] or cf. Eq. (24) in [16]):
δkpN(t1) = k ×(cid:16)(cid:16)c−1 ∆ xpN (t1) − R−1 ∆xpN (t1)(cid:17) × k(cid:17) = XA
pN(t1) = − (1 + γ)
k ×(cid:16)rA
δkA
G
c2
MA
rA
1
1 + rA
rA
0 rA
.
1 (cid:17)
0 × rA
0 · rA
1
δkA
pN(t1) ,
(10)
The sum in (10) runs over the bodies A of solar system. We also note the absolute value of
monopole term from one body A:
δkA
= (1 + γ)
Here, the impact parameter dA can be computed as
(cid:12)(cid:12)(cid:12)
pN(t1)(cid:12)(cid:12)(cid:12)
1 × k(cid:17) + O(c−2) = k ×(cid:16)rA
dA = k ×(cid:16)rA
because of σ = k + O (c−2).
MA
rA
1
G
c2
(cid:12)(cid:12)(cid:12)
1 (cid:12)(cid:12)(cid:12)
rA
0 × rA
0 · rA
1 + rA
rA
0 rA
1
.
0 × k(cid:17) + O(c−2) ,
(11)
(12)
III. THE QUADRUPOLE LIGHT DEFLECTION FOR STARS AND QUASARS
Using the expression ∆ xQ (t1) given by Eq. (44) of [12] and inserting into Eq. (4) one
gets [13]
δσQ(t1) =XA
δσA
Q(t1) ,
δσA
Q(t1) =
1 + γ
2
G
c2 "α′
A
UA(t1)
c
+ β′
A
EA(t1)
c
+ γ ′
A
FA(t1)
c
+ δ′
A
VA(t1)
c
# ,
(13)
where the sum in (13) runs over the massive bodies A of solar system. The scalar functions
are
UA
c
EA
c
FA
c
VA
c
=
1
d3
A
1(cid:17)2
= (cid:16)rA
2 + 3
1
(rA
1(cid:17)3
1 )3
1 − (cid:16)σ · rA
σ · rA
rA
1(cid:17)2
− 3 (cid:16)σ · rA
1 )5
(rA
σ · rA
1 )5 ,
(rA
,
1
=−3 dA
1
1 )3 ,
(rA
=−
,
(14)
(15)
(16)
(17)
and the time-independent vectorial coefficients
α
′ k
A =−M A
′ k
A = 2M A
β
γ
′ k
A = M A
ij
ij σi σj dk
A
dA
ij σi dj
A dk
A
d2
A
A dj
di
A dk
A
d3
A
dj
A
dA − 2 M A
ij σi σk dj
A
dA − 4 M A
ij
+ 2 M A
kj
,
− M A
ij σi σj dk
A
dA
,
′ k
A =−2 M A
δ
ij σi σj σk + 2 M A
kj σj − 4 M A
ij σi dj
A dk
A
d2
A
.
A dj
A dk
di
A
d3
A
,
6
(18)
(19)
(20)
(21)
The quadrupole formula (13) is valid for sources at infinite distance from the observer. The
sum over A in (13) runs, in principle, over all bodies inside the solar system, but only the
giant planets contribute within the accuracy of 1 µas; the quadrupole effect of the Sun is
irrelevant for Gaia mission because of the 45 degrees observation angle of the Sun by Gaia.
The symmetric and tracefree quadrupole moment of an object A is defined in [10 -- 13] and
given by
M A
ij =ZA
d3x ρA(x) (cid:18)ri rj −
1
3
δij r2(cid:19) ,
ri = xi − xi
A ,
(22)
with the mass density ρA, and the integral is taken over the volume of body A; the Kro-
necker symbol δij = 1 for i = j and zero otherwise. For an axial symmetric body (this
approximation is sufficient for the giant planets and aimed accuracy of 1µas) one has (see
Eqs. (48) -- (53) of [12])
M A
ij = MA J A
2 P 2
A
1
3 R
1
0
0
0
1
0
0
0 − 2
RT ,
(23)
where R is the rotational matrix giving the orientation of the symmetry (rotational) axis
e3 of the massive body in the BCRS, MA is the mass of the massive body A, J A
is the
2
coefficient of the second zonal harmonic of the gravitational field, PA is the minimal radius
of a sphere containing the body A and whose center coincides with the center of mass of A
(for the giant planets PA is just the equatorial radius).
IV. APPROXIMATION OF THE QUADRUPOLE LIGHT DEFLECTION
FORMULA FOR STARS AND QUASARS
In this Section we will derive an approximation of (13) sufficient for the envisaged accuracy
of 1 µas. From (13) one obtains the estimate
δσQ≤
1 + γ
2
G
c2 XA "α′
A UA
c
A EA
c
+ β′
A FA
c
+ γ ′
A VA
c # .
+ δ′
(24)
7
Note that since δσQ is perpendicular to σ, the absolute value δσQ gives, in the adopted
post-Newtonian approximation, the change of the calculated or observed direction to star
or quasar due to the quadrupole light deflection. Now we will estimate the terms in (24).
For such an estimation we consider the case of an axial symmetric body because this ap-
proximation is sufficient for the giant planets and goal accuracy of 1 µas, i.e. we take the
quadrupole tensor in the form (23).
A. Estimate of the vectorial coefficients
In this Section we estimate the vectorial coefficients of the last three individual terms in
A and
(24). For estimating the maximal possible absolute value of the coefficients β′
δ′
A, we replace R by the unit matrix. Then, by inserting (23) into (18) - (21) we obtain
A, γ ′
A = −MAJ A
α′
2 P 2
A
1
dA" 1 − (σ · e3)2 − 4
(dA · e3)2
d2
A
! dA + 2 (dA · e3) e3
−2 (σ · e3) (dA · e3) σ# ,
A = −2 MA J A
β′
2 P 2
A
1
d2
A
(σ · e3) (dA · e3) dA ,
A = −MA J A
γ ′
2 P 2
A
1
d3
A h(dA · e3)2 dA − (σ · e3)2 d2
AdAi ,
A = 2 MA J A
δ′
2 P 2
A "(σ · e3)2 σ +
2
d2
A
(σ · e3) (dA · e3) dA − (σ · e3) e3# ,
(25)
(26)
(27)
(28)
where e3 is the unit direction along the axis of symmetry (rotation). Here, (σ · e3) and
(dA · e3) are the projections of the vectors σ and dA, respectively, on the axis of symmetry.
With the aid of (26) - (28) we can explicitly determine the maximal absolute values of the
last three individual terms in (24):
J A
J A
2 (cid:12)(cid:12)(cid:12)
A≤ 2 MA (cid:12)(cid:12)(cid:12)
β′
2 (cid:12)(cid:12)(cid:12)
A = MA (cid:12)(cid:12)(cid:12)
γ ′
2 (cid:12)(cid:12)(cid:12)
A≤ 2 MA(cid:12)(cid:12)(cid:12)
δ′
J A
dA · e3
P 2
A ,
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A σ · e3
(dA · e3)2
A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤ MA (cid:12)(cid:12)(cid:12)
A (σ · e3)2 + 2(σ · e3) (dA · e3)
dA ≤ MA (cid:12)(cid:12)(cid:12)
− (σ · e3)2(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
d2
A
dA
P 2
P 2
J A
P 2
A ,
2 (cid:12)(cid:12)(cid:12)
+ (σ · e3)! ≤ 6 MA(cid:12)(cid:12)(cid:12)
(29)
(30)
P 2
A , (31)
J A
2 (cid:12)(cid:12)(cid:12)
(32)
where for the estimates (29) and (31) we have taken into account that
σ · e3 dA · e3
dA ≤
1
2
,
valid due to σ · dA = 0.
Parameter
Jupiter
Saturn
Uranus
Neptune
GMA/c2 [m]
J A
2 [10−3]
PA [106 m]
rmin
oA [1012 m]
GMA J A
2 P 2
A/c2 [1015 m3]
1.40987
0.42215
0.064473
0.076067
14.697
71.492
0.59
16.331
60.268
1.20
3.516
25.559
2.59
3.538
24.764
4.31
0.106
0.025
0.000148
0.000165
TABLE I: Numerical parameters of the giant planets taken from [17, 18].
B. Estimate of the scalar functions
Furthermore, from (15) - (17) we deduce the estimates
1(cid:17)2
+ 3 (cid:16)rA
1 )5
(rA
= 4
1
1 )3 ,
(rA
3 r1
1 )5 ≤ 3
(rA
1
1 )3 ,
(rA
1(cid:17)2
c ≤ (cid:16)rA
EA
FA
c ≤ dA
VA
c
1
1 )3 .
(rA
=
By inserting these estimates (29) - (31) and (33) - (35) into (24) we get
G
c2 β′
A EA
c ≤ 4
G
c2 γ ′
A FA
c ≤ 3
G
c2 δ′
A VA
c ≤ 6
G
G
c2 MA (cid:12)(cid:12)(cid:12)
c2 MA (cid:12)(cid:12)(cid:12)
c2 MA (cid:12)(cid:12)(cid:12)
G
J A
P 2
A
1
(rA min
1
)3 ,
P 2
A
1
(rA min
1
)3 ,
P 2
A
1
(rA min
1
)3 .
J A
J A
2 (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
8
(33)
(34)
(35)
(36)
(37)
(38)
The quantity rA min
1
represents the minimal distance between the object A and the observer.
C. Collection of all terms
Table I summarizes physical parameters of the giant planets. In this Table and in the
following discussions we use values of rA min
computed under assumption that the observer
is within a few million kilometers from the Earth's orbit. From the values given in Table I
1
and (36) - (38) we deduce (for these estimates γ = 1)
G
c2 "β′
A EA
c
A FA
c
+ γ ′
A VA
+ δ′
c #≤ 1.61 × 10−9 µas
≤ 4.52 × 10−11 µas
≤ 2.64 × 10−14 µas
≤ 7.66 × 10−15 µas
9
(39)
for Jupiter ,
for Saturn ,
for Uranus ,
for Neptune .
Obviously, by comparing the estimates given in (39) with the envisaged accuracy of 1µas we
EA and γ ′
VA, can
can conclude that these last three terms in (13), i.e. β′
A
A
safely be neglected. Accordingly, for Gaia mission, the simplified quadrupole light deflection
for stars and quasars valid on microarcsecond level of accuracy, reads
FA and δ′
A
δσQ =
1 + γ
2
G
c2 XA
α′
A
UA
c
,
(40)
with UA given by (14) and α′
A given by (18).
V. AN UPPER ESTIMATE OF THE QUADRUPOLE LIGHT DEFLECTION
FOR STARS AND QUASARS
The simplified expression of quadrupole light deflection (40) it still complicated. In order
to avoid evaluation of this term for each object in the data reduction, a simple criterion
is needed which allows one, with a few additional arithmetical operations, to judge if the
quadrupole light deflection should be computed for a given source and for a given accuracy.
To deduce such a criterion, we first evaluate the absolute value of the vectorial coefficient
(25),
A = MA (cid:12)(cid:12)(cid:12)
α′
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A(cid:16)1 − (σ · e3)2(cid:17) ,
(41)
which yields for the absolute value of light deflection angle caused by the quadrupole field
of an massive object A
δσA
Q =
1 + γ
GMA
2
c2
P 2
A(cid:16)1 − (σ · e3)2(cid:17) 1
d3
A
2 + 3
1
1(cid:17)3
1 )3
1 − (cid:16)σ · rA
σ · rA
rA
(rA
.
(42)
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
A comparison of (42) with the absolute value of spherically symmetric part given in (7) and
taking into account the fact
1(cid:17)3
1 − (cid:16)σ · rA
σ · rA
1 )3 + 2 ≤
rA
(rA
1
3
9
4 1 +
1
σ · rA
1 ! ,
rA
we obtain the criterion
δσA
Q ≤
9
8
P 2
A
d2
A (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12) (cid:16)1 − (σ · e3)2(cid:17) δσA
pN .
(43)
(44)
Due to 1 ≥ (σ · e3)2, the estimate (44) can be further approximated by
δσA
Q(cid:12)(cid:12)(cid:12) ≤
9
8 (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A
d2
A (cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
δσA
.
pN(cid:12)(cid:12)(cid:12)
This criterion relates the quadrupole light deflection for stars and quasars to the simpler
case of spherically symmetric part given in (7). It is recommended for Gaia to use (45) as
a criterion if the quadrupole light deflection has to be computed for a given star or quasar.
10
(45)
Eq. (42) can be used to estimate (cid:12)(cid:12)(cid:12)
Q(cid:12)(cid:12)(cid:12)
2 (1 + γ) GMA
δσA
directly:
(cid:12)(cid:12)(cid:12)
δσA
Q(cid:12)(cid:12)(cid:12) ≤
c2
where we have used 2 + 3 cos α − cos3 α ≤ 4. The estimate (46) coincides with [10] (see
Eq. (41) and the sentence below in that reference).
P 2
A
d3
A (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12) ≤
2 (1 + γ) GMA
c2 PA
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
,
(46)
VI. THE QUADRUPOLE LIGHT DEFLECTION FOR SOLAR SYSTEM
OBJECTS
The quadrupole light deflection for solar system objects δkQ is defined by Eqs. (36) -- (47)
and (69) of [12]. Using Eq. (9) it can be written as [13]:
δkQ =XA
δkA
Q ,
δkA
Q =
1 + γ
2
G
c2 "α′′
AAA(t1)
c
+ β′′
A BA(t1)
c
+ γ ′′
A CA(t1)
c
+ δ′′
A DA(t1)
c
# ,
(47)
where the sum in (47) runs over the massive bodies A of solar system. The scalar functions
are
dA
1 )3
(rA
1 − k · rA
2 rA
1 )2 ,
(rA
1 − k · rA
1
,
1
0
1
0
1
1
=
=
rA
0
1
rA
1
1
dA
AA
c
R 1
R k · rA
0 )3 −
(rA
R 1
0 )3 −
(rA
R k · rA
1
0 −
=−
rA
rA
0 + k · rA
1 + k · rA
rA
1 ! +
0 −
rA
0 − k · rA
rA
1 − k · rA
1(cid:17)2
1(cid:17)2
− 3 (cid:16)k · rA
1 )3 ! + (cid:16)rA
k · rA
1 )5
(rA
(rA
k · rA
1 )3! − 3 dA
1
1 )5 ,
(rA
(rA
k · rA
1 ! −
1
1 )3 ,
rA
(rA
and the time-independent vectorial coefficients are
BA
c
CA
c
DA
c
1
d2
A
dA
=
0
1
1
α′′ k
A = −M A
ij ki kj dk
A
dA
+ 2 M A
kj
β ′′ k
A = 2M A
ij ki dj
A dk
A
d2
A
,
dj
A
dA − 2 M A
ij ki kk dj
A
dA − 4 M A
ij
A dj
di
A dk
d3
A
,
(48)
(49)
(50)
(51)
(52)
(53)
A = M A
γ ′′ k
ij
A dj
A dk
di
A
d3
A
− M A
ij ki kj dk
A
dA
,
δ′′ k
A = −2 M A
ij ki kj kk + 2 M A
kj kj − 4 M A
ij ki dj
A dk
A
d2
A
11
(54)
(55)
.
In the following we will investigate how (47) can be simplified for a goal accuracy of 1 µas
and taking into account that in the case of Gaia the observer is situated within a few million
kilometers from the Earth's orbit.
VII. APPROXIMATION OF QUADRUPOLE LIGHT DEFLECTION FOR
SOLAR SYSTEM OBJECTS
To determine the magnitude of the individual terms in (47) we first notice the estimate
δkQ≤
1 + γ
2
G
c2 XA α′′
A AA
c
+ β′′
A BA
c
+ γ ′′
A CA
c
c ! .
A DA
+ δ′′
(56)
Since δkQ is perpendicular to k the absolute value δkQ(t1) gives, in the adopted post-
Newtonian approximation, the change of the calculated or observed direction to a solar
system object due to the quadrupole light deflection.
A. Estimate of the vectorial coefficients
In order to estimate the maximal value of vectorial coefficients we make use of the diago-
nalized form of quadrupole moment given in (23), which yields for the vectorial coefficients
(52) - (55)
A = −MA J A
α′′
2 P 2
A
1
d A" 1 − (k · e3)2 − 4
(dA · e3)2
d2
A
(57)
(58)
(59)
A = −2 MA J A
β′′
+2 (dA · e3) e3 − 2 (k · e3) (dA · e3) k +
A (cid:20)(k · e3) (dA · e3) dA −
A h(dA · e3)2 dA − (k · e3)2 d2
2 P 2
A
2 P 2
A
1
d3
1
d2
1
3
A = −MA J A
γ ′′
2
3
! dA
(k · dA) k#,
(k · dA) dA(cid:21) ,
A dAi ,
δ′′ = 2 MA J A
2 P 2
A "(k · e3)2 k +
2
d2
A
(k · e3) (dA · e3) dA − (k · e3) e3 −
2
3
1
d2
A
(k · dA) dA# ,
(60)
where k· e3 and dA · e3 are the projections of the vectors k and dA, respectively, on the axis
of symmetry.
From (58) - (60) we deduce the following absolute values for the last three vectorial
coefficients,
12
P 2
A ,
+
dA
1
3
P 2
P 2
d2
A
J A
2 (cid:12)(cid:12)(cid:12)
k · dA
dA ! ≤ MA (cid:12)(cid:12)(cid:12)
≤ MA (cid:12)(cid:12)(cid:12)
A k · e3 dA · e3
dA · e32
A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
− k · e32(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1
A(cid:20) 4
A (cid:16)3 (k · e3)3 − (k · e3)(cid:17) (k · dA) (dA · e3)
d2
(k · dA)2 (k · e3)2 − (k · e3)4 + (k · e3)2 +
−
P 2
A ,
3
1
d2
A
2 (cid:12)(cid:12)(cid:12)
P 2
A ,
J A
P 2
4
3
J A
J A
2 (cid:12)(cid:12)(cid:12)
A ≤ 2 MA (cid:12)(cid:12)(cid:12)
β′′
A = MA (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
γ ′′
2 (cid:12)(cid:12)(cid:12)
A = 2 MA (cid:12)(cid:12)(cid:12)
δ′′
2 (cid:12)(cid:12)(cid:12)
≤ MA (cid:12)(cid:12)(cid:12)
J A
J A
where for the estimates (61) and (63) we have taken into account that
k · e3 dA · e3
dA ≤
(k · e3)2 − (k · e3)4 ≤
1
2
1
4
,
.
(61)
(62)
4
9
1
d2
A
(k · dA)2(cid:21)1/2
(63)
(64)
(65)
The first estimate uses the fact that in post-Newtonian order k · dA = 0.
B. Estimate of the scalar funtions
In the following we estimate the magnitude of the scalar functions (49) - (51).
The coefficient given in (49) can be written as follows,
1. Estimate of BA
=
BA
c
0 · rA
Inserting the definition of vector k yields
1
q(rA
1
1 )2 − 2 rA
0 )2 + (rA
k · rA
0 )3 −
(rA
0
1(cid:17)2
1 )3 ! + (cid:16)rA
k · rA
(rA
1
1(cid:17)2
− 3(cid:16)k · rA
1 )5
(rA
.
(66)
1 rA
0
1 )2 cos α +
(rA
rA
1
0 )2 cos α −
(rA
1
0 −
rA
1
rA
1 !
BA
c
=
,
(rA
(rA
1
1 )2 − 2 rA
0 · rA
1(cid:17)2
− 3(cid:16)k · rA
1 )5
0 )2 + (rA
1(cid:17)2
+(cid:16)rA
rA
1 · rA
rA
1 rA
x2 + y2 − 2 x y cos α x
1
0
0
where cos α =
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
. By means of the inequality (with x = rA
0 , y = rA
1 )
y2 cos α +
y
x2 cos α −
1
x −
x + y
x2 y2
≤
1
y!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(67)
(68)
valid for any x ≥ 0 and y ≥ 0, we obtain the estimate
1
BA
c ≤
rA
0 + rA
0 )2 (rA
(rA
1
1 )2 +
4
1 )3 ≤
(rA
dA (rA
1 )2 +
1
A rA
d2
1
+
4
1 )3 .
(rA
The coefficient given in (50) can be written as follows,
2. Estimate of CA
CA
c
=
Since
dA
(rA
0 )3 (rA
1 )3
1(cid:17)3
(cid:16)rA
0 )2 + (rA
0(cid:17)3
−(cid:16)rA
1 )2 − 2 rA
0 · rA
1
q(rA
− 3 dA
1
k · rA
1 )5 .
(rA
q(rA
1
≤
1
q(rA
1 )2 − 2 rA
0 )2 + (rA
0 · rA
0(cid:17)3
1(cid:17)3
(cid:16)rA
−(cid:16)rA
0 − rA
rA
1
0 )3 (rA
1 )3
(rA
dA
1
0 − rA
1 )2
,
+ 3
dA
1 )4 .
(rA
we find for the absolute value
CA
c ≤
By means of the inequality
that is valid for any x and y, we obtain the estimate
x3 − y3
x − y ≤
3
2 (cid:16)x2 + y2(cid:17)
CA
c ≤
3
2
0(cid:17)2
1(cid:17)2
dA (cid:16)rA
+(cid:16)rA
1 )3 + 3
0 )3 (rA
(rA
dA
1 )4 ≤
(rA
3
2
1
1 )3 +
(rA
3
2
1
A rA
d2
1
+ 3
dA
1 )4 .
(rA
The coefficient given in (51) can be written as follows,
3. Estimate of DA
DA
c
= −
1
d2
A
1
0 )2 + (rA
1 )2 − 2 rA
0 · rA
1
q(rA
Inserting the definition of vector k yields
k · rA
0 −
rA
0
1
k · rA
1 ! −
rA
1
1 )3 .
(rA
= −
With the aid of the inequality
DA
c
1
d2
A
0 cos α + rA
rA
0 )2 + (rA
(rA
1 cos α − rA
1 )2 − 2 rA
0 − rA
0 · rA
1
1
1
1 )3 .
(rA
−
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
x cos α + y cos α − x − y
x2 + y2 − 2 x y cos α (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
2
x + y
≤
valid for x ≥ 0 and y ≥ 0, we obtain the estimate
1
+
rA
1
DA
c ≤ 2
1
d2
A
1
1 )3 .
(rA
13
(69)
(70)
(71)
(72)
(73)
(74)
(75)
(76)
(77)
(78)
C. Collection of all terms
Altogether, by inserting the estimates of vectorial coefficients, (61) - (63), and the scalar
coefficients (69), (74), (78) into (56) yields
14
c
G
G
J A
c2 β′′
A BA
2 (cid:12)(cid:12)(cid:12)
c2 MA (cid:12)(cid:12)(cid:12)
≤
c2 MA (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
J A
≤
G
P 2
A CA
+ γ ′′
A "9
A "9
c
1
A rA
d2
1
1
P 2
P 2
2
2
+ δ′′
+
c !
A DA
1 )2 +
1
dA (rA
1
+
A rA min
1
PA (rA min
1
13
2
1
1 )3 + 3
(rA
)2 +
19
2
(rA min
1
dA
(rA
1
1 )4#
)3# ,
(79)
where we have used that PA ≤ dA ≤ rA
1 . Note, in the last line of (79) the first term in
the brackets is at least by a factor of ∼ 104 larger than the other two terms. Using the
parameters given in Table I we obtain for the giant planets (γ can be safely set to unity for
these estimates)
G
c2 β′′
A BA
c
+ γ ′′
A CA
c
+ δ′′
A DA
c ! ≤ 3.26 × 10−2µas
≤ 5.32 × 10−3µas
≤ 8.11 × 10−4µas
≤ 5.79 × 10−5µas
for Jupiter ,
for Saturn ,
for Uranus ,
for Neptune .
(80)
In view of these estimates, for the envisaged accuracy of 1 µas the quadrupole light deflection
(47) for sources in the solar system can be approximated by
A AA
α′′
c
δkQ =
1 + γ
(81)
G
2
,
c2 XA
with AA given by (48) and α′′
A given by (52).
VIII. AN UPPER ESTIMATE OF THE QUADRUPOLE LIGHT DEFLECTION
FOR SOLAR SYSTEM OBJECTS
The simplified expression of quadrupole light deflection for solar system objects (81) it
still complicated.
In order to avoid evaluation of this term for each object in the data
reduction, a simple criterion is needed which allows one, with a few additional arithmetical
operations, to judge if the quadrupole light deflection should be computed for a given source
and for a given accuracy. To deduce such a criterion, we first consider the absolute value of
the vectorial coefficient (57) given by
so that an estimate of the absolute value of one term in (81) is
A = MA (cid:12)(cid:12)(cid:12)
α′′
c2 MA (cid:12)(cid:12)(cid:12)
G
J A
2 (cid:12)(cid:12)(cid:12)
2 (cid:12)(cid:12)(cid:12)
J A
δkA
Q =
P 2
A(cid:16)1 − (σ · e3)2(cid:17) ,
P 2
A (cid:16)1 − (σ · e3)2(cid:17) AA
c
(82)
(83)
,
where the scalar coefficient AA is given in (48). According to Eq. (12) we have
and we obtain
0(cid:17)2
(cid:16)k × rA
1(cid:17)2
=(cid:16)k × rA
= d2
A + O(cid:16)c−2(cid:17) ,
=
1
d3
A
AA
c
1
R 1
rA
0(cid:17)2
0 + k · rA
−
1(cid:17)(cid:16)rA
1 − k · rA
R , and collecting all terms together we obtain
0 (cid:16)rA
A (cid:16)2 rA
1
1 )3
(rA
1
d3
1
rA
+
1 (cid:16)rA
1 + k · rA
1(cid:17)2
1 + k · rA
.
1(cid:17)2!
Using k = R
15
(84)
(85)
AA(t1)
c
=
1
d3
A
1
R3 (1 − cos α)2 (cid:18)2 (cid:16)rA
0(cid:17)3
1(cid:17)2
+(cid:16)rA
rA
0(cid:17)2
0 + 2 (cid:16)rA
rA
0(cid:17)3
1 +(cid:16)rA
cos α(cid:19) . (86)
By means of the inequality (see Appendix A for a proof)
(1 − cos α)2 2x3 + xy2 + 2x2y + x3 cos α
(x2 + y2 − 2xy cos α)3/2 ≤ 3
valid for any x ≥ 0 and y ≥ 0 and with the aid of
x
(x2 + y2 − 2xy cos α)1/2
sin2 α
1 + cos α
,
(87)
we obtain the estimate,
rA
0
R
sin α =
dA
rA
1
,
AA
c ≤ 3
1
rA
1
1
d2
A
sin α
1 + cos α
,
and, therefore, with the aid of (81) and (82) we achieve
δkA
Q≤
3 (1 + γ)
GMA
2
c2
1
d2
A
This result can be related to the spherically symmetric part, given in (11). By comparison
between (90) and (11) we obtain the criterion
1
rA
1 (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A (cid:16)1 − (σ · e3)2(cid:17)
sin α
1 + cos α
.
(88)
(89)
(90)
(91)
(92)
(cid:12)(cid:12)(cid:12)
δkA
Q(cid:12)(cid:12)(cid:12)≤
3
2
P 2
A
d2
A (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12) (cid:16)1 − (σ · e3)2(cid:17) (cid:12)(cid:12)(cid:12)
δkA
.
pN(cid:12)(cid:12)(cid:12)
Due to 1 ≥ (σ · e3)2, the estimate (91) can be further approximated by
δkA
Q ≤
3
2 (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A δkA
A
d2
pN .
This criterion relates the quadrupole light deflection of sources in the solar system to the
simpler case of spherically symmetric part. For Gaia it is recommended to use (92) as
a criterion if the quadrupole light deflection has to be calculated for a given solar system
object. The estimate of the monopole light deflection for solar system objects can be written
as follows:
16
2 (1 + γ) GMA
2 (1 + γ) GMA
≤
(cid:12)(cid:12)(cid:12)
δkA
pN(cid:12)(cid:12)(cid:12)≤
c2 dA
which can be used in the case when δkA
pN is not available; the proof of (93) is straight-
forward by means of (11). From (83) and (86) one can directly see that (for a proof see
Appendix (B))
c2 PA
,
(93)
(cid:12)(cid:12)(cid:12)
δkA
Q(cid:12)(cid:12)(cid:12)≤ 2 (1 + γ)
where we have used PA ≤ dA.
GMA
c2
P 2
A
A J2 ≤ 2 (1 + γ)
d3
GMA
c2 PA J2 ,
(94)
IX. SHAPIRO EFFECT FOR SOLAR SYSTEM OBJECTS
In this Section,
for reasons of completeness, the known analytical expressions of
quadrupole Shapiro effect will be reconsidered, which is also of practical interest for as-
trometric missions in nearest future; e.g. BepiColumbo or Juno require modelling of the
light travel time at the level of millimeters. From Eq. (1) we obtain for the Shapiro effect
in post-Newtonian order the expression
where cτ = c(t1 − t0) and
cτ = R + cXi
δτi + O(cid:16)c−4(cid:17) ,
c δτi = −k · ∆xi(t1) .
In order to show (95) and (96) we have used (cf. Eq. (23) of [10])
σ = k − R−1 k × Xi
∆xi(t1) × k! + O(cid:16)c−4(cid:17) .
(95)
(96)
(97)
The sum in (95) runs over the terms of the metric caused by the massive body (e.g. spherical
symmetric term, quadrupole term and higher multipole terms etc.). Here, for our purposes
it will be sufficient to consider the spherical symmetric and quadrupole part. The spherical
symmetric term of Shapiro effect is given by
c δτpN =−k · ∆pN x(t1) = XA
c δτ A
pN ,
c δτ A
pN = (1 + γ)
G
c2 MA log
rA
0 + rA
rA
0 + rA
1 + R
1 − R
,
(98)
where the sum runs over all massive bodies A under consideration. The quadrupole term of
Shapiro effect is given by
c δτQ = −k · ∆Q x(t1) = XA
c δτ A
Q .
(99)
The expression ∆Q x(t1) has been given in [12]. Accordingly, the Shapiro effect for
quadrupole gravitational fields of one massive solar system body A is given by
17
c δτ A
Q =
1 + γ
2
G
c2 (δA VA + βA EA + γA FA) ,
with the scalar functions
1
k · rA
1 )3 ,
(rA
1
(rA
0
k · rA
EA =
0 )3 −
(rA
FA = dA 1
VA =−
1
d2
0 )3 −
(rA
A k · rA
0 −
rA
0
1 )3! ,
k · rA
1 ! ,
rA
1
and the time-independent scalar coefficients are
di
A
dA
dj
A
dA
,
βA = M A
γA = 2 M A
ij
ij ki kj − M A
ij ki dj
A
dA
,
δA = M A
ij ki kj + 2 M A
ij
di
A
dA
dj
A
dA
.
(100)
(101)
(102)
(103)
(104)
(105)
(106)
Note, the impact vector can be computet by means of Eq. (12). In Appendix C we show
the following estimates:
G MA
G
c2 δA VA ≤
G
c2 βA EA + γA FA ≤
(cid:12)(cid:12)(cid:12)
c2
J A
P 2
A
d2
A
G MA
2 (cid:12)(cid:12)(cid:12)
c2
,
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A
0 )2 +
(rA
P 2
A
(rA
1 )2! .
(107)
(108)
0 , rA
1 , we conclude the inequality
These estimates imply that (100) cannot be simplified for the general case. Furthermore,
from these relations, because of PA ≤ dA, rA
Q (cid:12)(cid:12)(cid:12)≤ 3 (cid:12)(cid:12)(cid:12)
which represents a strict upper bound of quadrupole Shapiro effect and improves the estimate
given in Eq. (47) in [10]. This estimate implies that for quadrupole light deflection there is
a maximal numerical value which depends only on physical parameters of the massive body,
but not on distance dA. Numerical values of the estimate (109) for the giant planets are
given in Table II.
2 (cid:12)(cid:12)(cid:12)
G MA
c δτ A
(109)
J A
(cid:12)(cid:12)(cid:12)
c2
,
X. NUMERICAL TESTS
The obtained simplified formulas given by Eq. (40) for stars/quasars and by Eq. (81) for
solar system objects and the a priori criteria given for these objects by Eq. (45) and (92),
Parameter
Sun
Jupiter
Saturn
Uranus
Neptune
18
G MA
c2
[mm]
0.89
62.16
20.68
0.68
0.81
3 (cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
TABLE II: Numerical values of estimate (109). For the Sun a value J ⊙
adopted [19]; GM⊙/c2 = 1476 m.
2 = 2 × 10−7 has been
respectively, have been incorporated into the current reference C implementation of GREM
as described by Klioner & Blankenburg [13] and Klioner [14].
Numerical experiments with the C implementation have confirmed the correctness and
the efficiency of the estimates and criteria. In the experiments we used about 108 objects
(both randomly distributed over the sky and specially generated to give grazing rays to the
giant planets). The results can be summarized as follows:
I. Stars and quasars.
-- The maximal difference between the full quadrupole deflection formula (13) and
the simplified one (40) amounts to 1.1× 10−10 µas in good agreements with (39).
Hence, the actual values of the neglected terms are in case of Gaia about 10 -- 15
times less than given by (39).
-- The upper estimate (45) holds and is attainable for randomly distributed sources.
-- The mean value of the ratio between the actual value of the quadrupole light
deflection and its upper estimate (45) amounts to 0.48 for randomly distributed
sources that indicates the high numerical efficiency of the estimate.
II. Solar system objects.
-- The maximal difference between the full quadrupole deflection formula (47) and
the simplified one (81) amounts to 0.0017 µas in good agreements with (80).
Therefore, the actual values of the neglected terms are again about 10 -- 15 times
less than given by (80).
-- The upper estimate (92) holds and is attainable for randomly distributed sources.
-- The mean value of the ratio between the actual value of the quadrupole light
deflection and its upper estimate (92) amounts to 0.40 for randomly distributed
sources that again indicates the high numerical efficiency of the estimate.
III. Quadrupole Shapiro effect.
-- A strict upper estimate of quadrupole Shapiro effect is given in Eq. (109).
Implementation of the criteria (45) and (92) has allowed to significantly reduce the num-
ber of "false alarms" (cases for which the full quadrupole deflection has been computed and
turned out to be much smaller than the requested goal accuracy). The "false alarms" were
caused by the use of a more primitive ad hoc criteria implemented in the C code of GREM
previously. This, in turn, slightly increases the performance of the GREM implementation.
19
Let us summarize the results of this report.
XI. SUMMARY
1. Quadrupole light deflection (13) for stars and quasars is approximated by (40).
2. Eqs. (45) and (46) can be used as an a priori criterion if the quadrupole light deflection
(40) has to be computed for a given source.
3. Quadrupole light deflection (47) for solar system sources is approximated by (81).
4. Eqs. (92) and (94) can be used as an a priori criterion if the quadrupole light deflection
(81) has to be computed for a given solar system object.
5. A strict criterion of quadrupole Shapiro effect has been given in Eq. (109), which
improves the estimate given in Eq. (47) of [10].
Our investigations, i.e. the simplified quadrupole formulas and the criteria, provide a
highly time-efficient tool for data reduction on microarcsecond level of accuracy for Gaia
mission.
[1] O.S. Ivanitskaya, Lorentz Basis and Gravitational Effects in Einstein's Theory of Gravitation
(1979) Nauka i Tekhnika, Minsk (in Russian).
[2] Chr. Le Poncin-Lafitte, P. Teyssandier, Phys. Rev. D 77 (2008) 044029.
[3] S.M. Kopeikin, J. Math. Phys. 38 (1997) 2587.
[4] S.M. Kopeikin, P. Korobkov, A. Polnarev, Class. Quantum Grav. 23 (2006) 4299.
[5] S.M. Kopeikin, V.V. Makarov, Phys. Rev. D 75 (2007) 062002.
[6] R. Epstein, I.I. Shapiro, Phys. Rev. D 22 (1980) 2947.
[7] G.W. Richter, A. Matzner, Phys. Rev. D 26 (1982) 1219.
[8] G.W. Richter, A. Matzner, Phys. Rev. D 26 (1982) 2549.
[9] S.A. Cowling, Mon. Not. R. astr. Soc. 209 (1984) 415.
[10] S.A. Klioner, Sov. Astron., 35 (1991) 523.
[11] S.A. Klioner, S.M. Kopejkin, AJ 104 (1992) 897.
[12] S.A. Klioner, AJ 125 (2003) 1580.
[13] S.A. Klioner,
(2003) Technical
R. Blankenburg
on
tion of
http://www.rssd.esa.int/llink/livelink.
the Gaia relativistic model,
available
report
implementa-
from the Gaia document archive
the
[14] S.A. Klioner
(2003) Technical report on the implementation of
the Gaia relativis-
tic model, Amendment for version 1.0g, available from the Gaia document archive
http://www.rssd.esa.int/llink/livelink.
[15] G.W. Richter, A. Matzner, Phys. Rev. D 28 (1982) 3007.
[16] S.A. Klioner, S. Zschocke, Class. Quantum Grav. 27 (2001) 075015.
[17] IERS Conventions (2003). D.D. McCarthy, G. Petit, IERS Conventions (IERS Technical Note
32) Frankfurt am Main: Verlag des Bundesamts fur Kartographie und Geodasie, 2004, 127
pp.
[18] P.R. Weissman, L.-A. McFadden, T.V. Johnson, Encyclopedia of the Solar System, (San Diego:
Academic) eds. 1999.
[19] A. Fienga, H. Manche, J. Laskar, M. Gastineau, A new numerical planetary ephemeris, A&A
477 (2008) 315.
20
Appendix A: Proof of Eq. (87)
In this Appendix we prove the inequality (87). The latter can be rewritten as
(1 − cos α)2 (2x3 + x y2 + 2x2 y + x3 cos α)≤ 3 x (x2 + y2 − 2xy cos α) (1 − cos α) .
Denoting z = x/y we obtain the relation
f1 ≡−1 − 2 z2 + 2 z − cos α + 4 z cos α − cos2 α − z2 cos α ≤ 0 .
21
(A1)
(A2)
To prove the inequality (A2) it is sufficient to investigate the values of f at extrema and at
the boundaries given by z ≥ 0 and 0 ≤ α ≤ π. To determine the extrema of this function
we set the first derivatives zero,
f1 , z = −4 z + 2 + 4 cos α − 2 z cos α = 0 ,
f1 , α = sin α (1 − 4 z + 2 cos α + z2) = 0 .
The only solutions of the coupled system (A3), (A4) are
P1 = (z = 0, α =
2
3
π) , P2 = (z = 1, α = 0) .
The values of f1 at these points are
At the boundaries we get
f1(P1) = −
3
4
,
f1(P2) = 0 .
f1(z = 0) =−1 − cos α − cos2 α < 0 ,
f1(z → ∞) =−z2(2 + cos α) < 0 ,
f1(α = 0) =−3(1 − z)2 ≤ 0 ,
f1(α = π) =−z(z + 2) ≤ 0 .
(A3)
(A4)
(A5)
(A6)
(A7)
(A8)
(A9)
(A10)
From the results (A5) - (A10) we conclude the validity of the inequality (A2) and (A1).
Appendix B: Proof of Eq. (94)
From (83) it is clear that (94) is true if for AA defined by (86) one has
AA(t1)
c
≤
4
d3
A
.
To prove this it is sufficient to demonstrate that
1
R3 (1 − cos α)2 (cid:18)2 (cid:16)rA
0(cid:17)3
1(cid:17)2
+(cid:16)rA
rA
0(cid:17)2
0 + 2 (cid:16)rA
rA
0(cid:17)3
1 +(cid:16)rA
cos α(cid:19)≤ 4 ,
(B1)
(B2)
or introducing again z = rA
1
0 /rA
(1 − cos α)2 z ((1 + z)2 + z2 (1 + cos α))
4 (1 + z2 − 2z cos α)3/2
f2 ≡
The derivatives of f2 with respect to α and z vanish simultaneously only for α = 0 which is
one of the boundaries. At the boundaries we get
22
(B3)
≤ 1 .
f2(z = 0) = 0 ,
1
4
f2(α = 0) = 0 ,
lim
z→∞
f2 =
f2(α = π) =
z
1 + z ≤ 1 .
(1 − cos α)2 (2 + cos α) ≤ 1 ,
From this we conclude that (B2) and, therefore, (B1) and (94) are valid.
Appendix C: Proof of inequalities (107) and (108)
First, we consider the inequality (107), which is given by
G
c2 δA VA =
≤
1
d2
1
d2
A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
M A
ij ki kj + 2 M A
ij
di
A
dA
M A
ij ki kj + 2 M A
ij
di
A
dA
1
k · rA
rA
1 !(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
0
dj
dj
A
dA! k · rA
0 −
rA
dA (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
,
where the inequality in (C2) is obvious. By inserting (23) into (C2) we obtain
G
c2 δA VA ≤
G MA
c2
Using the inequality
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
P 2
A
d2
A (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1 − (k · e3)2 − 2
(dA · e3)2
d2
A
.
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(B4)
(B5)
(B6)
(B7)
(C1)
(C2)
(C3)
(C4)
1 − (k · e3)2 − 2
(dA · e3)2
d2
A
≤ 1 ,
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
which can be shown by introducing spherical coordinates and taking into account that in
post-Newtonian order k and dA are perpendicular to each other (cf. proof of Eq. (C10)),
we find
G
c2 δA VA ≤
G MA
c2
P 2
A
d2
A
.
(C5)
Now, we consider the inequality (108). From the definitions of the functions (101) and (102)
and of the scalar coefficients (104) - (105), we obtain
G
c2 βA EA + γA FA
di
M A
A
dA
ij ki kj − M A
ij
f3
0 )2 +
(rA
f4
1 )2 ,
(rA
=(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
≤
dj
A
dA! k · rA
0 )3 −
(rA
0
1
k · rA
1 )3 ! + 2 M A
(rA
ij ki dj
A
dA dA
0 )3 −
(rA
dA
(rA
1 )3!(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(C6)
where the functions are given by
ij ki kj − M A
ij
ij ki kj − M A
ij
di
A
dA
di
A
dA
0
dj
A
rA
0
dA! k · rA
dA! k · rA
dj
A
rA
1
1
+ 2 M A
+ 2 M A
ij ki dj
A
dA
ij ki dj
A
dA
M A
M A
f3 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
f4 =(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
dA
rA
dA
rA
0 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
1 (cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
Now we consider function f3 (proof for function f4 is very similar). Inserting (23) into (C7)
yields
23
(C7)
(C8)
,
.
dA
rA
0 (cid:12)(cid:12)(cid:12)(cid:12)
1 )2! ,
P 2
A
(rA
≤ 1 ,
(C10)
(C11)
Using the inequality (see below)
f3 =
J A
G MA
2 (cid:12)(cid:12)(cid:12) (cid:12)(cid:12)(cid:12)(cid:12)
A (cid:12)(cid:12)(cid:12)
c2 P 2
(dA · e3)2
d2
A
h =(cid:12)(cid:12)(cid:12)(cid:12)
we find
(dA · e3)2
d2
A
− (k · e3)2! k · rA
0
0 − 2 (k · e3)
rA
(dA · e3)
dA
dA
rA
0 (cid:12)(cid:12)(cid:12)(cid:12)
.
(C9)
− (k · e3)2! k · rA
0
0 − 2 (k · e3)
rA
(dA · e3)
dA
G
c2 βA EA + γA FA ≤
G MA
c2
which is just relation (108).
P 2
A
0 )2 +
(rA
(cid:12)(cid:12)(cid:12)
J A
2 (cid:12)(cid:12)(cid:12)
In order to complete the proof of (C11), we still have to show relation (C10). For that
we introduce spherical coordinates as follows: dA = (0, 0, dA)T, k = (cos φ1 , sin φ1 , 0)T
and e3 = (sin θ cos φ2 , sin θ sin φ2 , cos θ)T, where we have taken into account that in post-
Newtonian order dA and k are perpendicular to each other. Accordingly, we have dA · e3 =
dA cos θ and k · e3 = sin θ (cos φ1 cos φ2 + sin φ1 sin φ2), and by inserting into Eq. (C10) we
obtain
h =(cid:12)(cid:12)(cid:12)(cid:16)cos2 θ − sin2 θ cos2 α(cid:17) cos Φ − 2 (cos θ sin θ cos α) sin Φ(cid:12)(cid:12)(cid:12)
k · rA
rA
0
0
= cos Φ and dA
rA
0
where we have used
= sin Φ, and the addition theorem of cosine and
sine function: cos φ1 cos φ2 + sin φ1 sin φ2 = cos α where α = φ1 − φ2. The expression in
(C12) can be estimated by A cos Φ + B sin Φ ≤ √A2 + B2; accordingly we obtain:
,
(C12)
h≤ cos2 θ + sin2 θ cos2 α ≤ cos2 θ + sin2 θ = 1 .
(C13)
Thus, we have shown relation (C10).
|
1811.09628 | 3 | 1811 | 2019-03-18T01:01:46 | Multiple Spiral Arms in Protoplanetary Disks: Linear Theory | [
"astro-ph.EP"
] | Recent observations of protoplanetary disks, as well as simulations of planet-disk interaction, have suggested that a single planet may excite multiple spiral arms in the disk, in contrast to the previous expectations based on linear theory (predicting a one-armed density wave). We re-assess the origin of multiple arms in the framework of linear theory, by solving for the global two-dimensional response of a non-barotropic disk to an orbiting planet. We show that the formation of a secondary arm in the inner disk, at about half of the orbital radius of the planet, is a robust prediction of linear theory. This arm becomes stronger than the primary spiral at several tenths of the orbital radius of the planet. Several additional, weaker spiral arms may also form in the inner disk. On the contrary, a secondary spiral arm is unlikely to form in the outer disk. Our linear calculations, fully accounting for the global behavior of both the phases and amplitudes of perturbations, generally support the recently proposed WKB phase argument for the secondary arm origin (as caused by the intricacy of constructive interference of azimuthal harmonics of the perturbation at different radii). We provide analytical arguments showing that the process of a single spiral wake splitting up into multiple arms is a generic linear outcome of wave propagation in differentially rotating disks. It is not unique to planet-driven waves and occurs also in linear calculations of spiral wakes freely propagating with no external torques. These results are relevant for understanding formation of multiple rings and gaps in protoplanetary disks. | astro-ph.EP | astro-ph |
Draft version March 19, 2019
Preprint typeset using LATEX style emulateapj v. 12/16/11
MULTIPLE SPIRAL ARMS IN PROTOPLANETARY DISKS: LINEAR THEORY
Ryan Miranda1,3 and Roman R. Rafikov1,2
Draft version March 19, 2019
ABSTRACT
Recent observations of protoplanetary disks, as well as simulations of planet-disk interaction, have
suggested that a single planet may excite multiple spiral arms in the disk, in contrast to the previous
expectations based on linear theory (predicting a one-armed density wave). We re-assess the origin of
multiple arms in the framework of linear theory, by solving for the global two-dimensional response
of a non-barotropic disk to an orbiting planet. We show that the formation of a secondary arm in
the inner disk, at about half of the orbital radius of the planet, is a robust prediction of linear theory.
This arm becomes stronger than the primary spiral at several tenths of the orbital radius of the
planet. Several additional, weaker spiral arms may also form in the inner disk. On the contrary, a
secondary spiral arm is unlikely to form in the outer disk. Our linear calculations, fully accounting
for the global behavior of both the phases and amplitudes of perturbations, generally support the
recently proposed WKB phase argument for the secondary arm origin (as caused by the intricacy of
constructive interference of azimuthal harmonics of the perturbation at different radii). We provide
analytical arguments showing that the process of a single spiral wake splitting up into multiple arms
is a generic linear outcome of wave propagation in differentially rotating disks. It is not unique to
planet-driven waves and occurs also in linear calculations of spiral wakes freely propagating with no
external torques. These results are relevant for understanding formation of multiple rings and gaps
in protoplanetary disks.
Subject headings: hydrodynamics -- protoplanetary disks -- planet -- disk interactions -- waves
1.
INTRODUCTION
The gravitational interaction of young planets with
their natal disks is known to produce spiral density
waves. Recent high-resolution direct imaging has re-
vealed spiral structures in several protoplanetary disks,
including MWC 758 (Grady et al. 2013; Benisty et al.
2015), HD 100453 (Wagner et al. 2015, 2018), and SAO
206462 (Muto et al. 2012; Garufi et al. 2013; Stolker et al.
2016; Maire et al. 2017), which may be produced by the
gravitational influence of planets or binary companions.
A remarkable feature of these disks is that they display
pairs of spiral arms separated by approximately 180◦,
which has not been expected.
Indeed, in the conventional picture of planet-disk in-
teractions, a planet is believed to give rise to only a sin-
gle spiral arm (Ogilvie & Lubow 2002; Rafikov 2002a).
Gravitational perturbations due to a planet excite many
wave modes in the disk, each described by an azimuthal
number m, and launched at a corresponding Lindblad
resonance -- locations interior and exterior to the or-
bit of the planet where orbital commensurabilities oc-
cur (Goldreich & Tremaine 1979). Ogilvie & Lubow
(2002) showed that the these modes interfere construc-
tively, leading to a characteristic one-armed spiral pat-
tern. In this framework, two planets would be required
to produce two arms (Benisty et al. 2015). However,
for pairs of arms with similar azimuthal separations to
be found in several unrelated protoplanetary disks would
1 Institute for Advanced Study, Einstein Drive, Princeton, NJ
08540
2 Centre for Mathematical Sciences, Department of Applied
Mathematics and Theoretical Physics, University of Cambridge,
Wilberforce Road, Cambridge CB3 0WA, UK
3 [email protected]
require fortuitous configurations of the orbital phases of
the planets in these systems.
Recent three-dimensional simulations of planet-disk in-
teractions have demonstrated that some of the observed
multiple spiral features can, in fact, be produced by a
single orbiting companion (Zhu et al. 2015; Dong et al.
2015b; Fung & Dong 2015; Dong et al. 2016a; Dong &
Fung 2017). Notably, the spirals seen in HD 100453 were
demonstrated to be consistent with the disturbances pro-
duced by the nearby M dwarf companion (Dong et al.
2016b; Wagner et al. 2018). A key finding of these nu-
merical studies is that a single planet can produce multi-
ple spirals, and so it is not necessary to invoke the pres-
ence of multiple planets to explain the appearance of
several spiral arms. In some cases, more than two spiral
arms are produced. The number of arms, as well as the
azimuthal separation of the two strongest spirals, were
found numerically to depend on planet mass (Zhu et al.
2015; Fung & Dong 2015; Bae & Zhu 2018b). It was also
shown (Bae et al. 2017) that multiple arms can be re-
lated to the formation of annular gaps in mm-size dust
distribution in protoplanetary disks.
At the same time, Arzamasskiy & Rafikov (2018) have
recently demonstrated numerically that formation of sec-
ondary spirals does not necessarily require the presence of
a planet (i.e., an orbiting point mass perturber) driving
density waves. In their case, density waves were driven
by an imposed boundary condition at the outer edge of
the simulation domain and then freely propagated in-
ward, without angular momentum injection by external
torques. Such passive propagation of the waves suffi-
ciently far into the inner disk was found to also naturally
result in the formation of a secondary arm.
Despite these numerical experiments, the origin of sec-
2
ondary spiral arms has remained elusive. Some nonlinear
mechanisms, such as mode coupling (Fung & Dong 2015;
Lee 2016) have been proposed to explain their features.
Recently, Bae & Zhu (2018a,b) argued that the forma-
tion of multiple spirals can be explained by radially-
dependent coherence of different azimuthal harmonics of
the perturbations driven by a planet, essentially by an
extension of the linear mode phase argument of Ogilvie
& Lubow (2002). In their work, the phases of the mul-
tiple crests of each mode were shown to constructively
interfere in different parts of the disk (at different az-
imuthal locations) as the wave propagates away from
the perturber. The different regions of interference are
identified as the primary arm, secondary arm, tertiary
arm, and so on. This argument was laid out in terms
of the local (WKB) approximation for mode phases and
essentially ignored the behavior of the mode amplitudes.
Nevertheless, these findings were corroborated by two-
dimensional numerical simulations, demonstrating this
idea to be a promising step towards understanding the
formation of multiple spirals. An important aspect of
the work of Bae & Zhu (2018a) is that the emergence of
multiple arms was understood, at least in part, within a
linear framework.
In this paper, we directly apply the linear theory of
density wave evolution to self-consistently compute the
full two-dimensional structure of the response of a thin,
locally isothermal disk to an orbital companion. By prop-
erly accounting for the global behavior of the mode am-
plitudes as well as their phases (i.e., going beyond the
WKB approximation), we show that multiple spiral arms
are robustly formed in the inner regions of protoplan-
etary disks; under certain circumstances they can also
appear in the outer disk. We characterize the morphol-
ogy of the spirals (e.g., their amplitudes, widths, and
arm-to-arm separations) and its dependence on the disk
properties -- its aspect ratio, as well as profiles of the
temperature and surface density.
The plan of this paper is as follows. In Section 2, we
describe our setup and the details of our calculations. In
Section 3, we present results on the formation of multi-
ple spirals by a planet, and characterize the properties
of the spirals and their dependence on the disk parame-
ters. In Section 4 we present calculations of the passive
propagation of a spiral wake in a perturber-free disk,
demonstrating that the emergence of a secondary spiral
is a generic property of wave propagation in differentially
rotating disks. In Section 5 we provide theoretical argu-
ments based on linear mode phases in order to under-
stand some key aspects of our calculations. We discuss
our results in Section 6, and conclude with a summary
of our main results in Section 7.
2. PROBLEM FRAMEWORK
We consider propagation of density waves in a two-
dimensional fluid disk around a star of mass M∗ in the
linear regime. The foundations of the mathematical
framework for studying this phenomenon were laid out
in Goldreich & Tremaine (1979), and we heavily borrow
from their results. We explore both the inhomogeneous
and homogeneous versions of the problem.
In the inhomogeneous case (§3), the wave is explicitly
driven by the gravitational potential of a planet of mass
Mp (cid:28) M∗ moving on a circular orbit with radius rp and
Keplerian frequency Ωp = (GM∗/r3
p)1/2. The torque due
to the planetary gravity both excites the wave in the first
place and modifies its subsequent propagation.
In the homogeneous case, the perturbation is imposed
at the edge of the disk with no external torques affect-
ing subsequent wave propagation (a setup analogous to
Arzamasskiy & Rafikov 2018). This regime is studied
using the same mathematical framework as the inhomo-
geneous case but with the planetary source terms set to
zero (§4).
2.1. Basic Setup
We consider a very general disk model in which the
entropy S ∝ ln(P/Σγ) is allowed to vary with radius
r. Here Σ is the disk surface density, P = Σc2
s /γ is
the (height-integrated) pressure, cs is an adiabatic sound
speed, and γ is the adiabatic index. We assume that in
the unperturbed disk,
cs(r) = hprpΩp
,
(1)
where hp is the disk aspect ratio, h(r) = H/r =
hp(r/rp)(1−q)/2, evaluated at rp, and H = cs/Ω is the
pressure scale height. The exponent q is the power law
index of the disk temperature, which is proportional to
c2
s . The disk surface density is
(cid:19)−q/2
(cid:18) r
rp
(cid:18) r
(cid:19)−p
rp
Σ(r) = Σp
,
(2)
where the value of Σp is arbitrary. As a result of spec-
ifying cs and Σ independently, the entropy S can vary
through the disk.
The orbital frequency of the disk fluid is modified from
the pure Keplerian value ΩK = (GM∗/r3)1/2 by the pres-
sure gradient:
Ω2 = Ω2
K +
1
rΣ
dP
dr
.
(3)
The response of the disk is sensitive to the small devi-
ations of Ω and the radial epicyclic frequency κ, given
by
κ2 =
2Ω
r
d
dr
(r2Ω),
(4)
from the Keplerian frequency ΩK.
2.2. Equations and Numerical Procedure
The surface density perturbation produced in re-
sponse to the gravitational potential of the planet (or
the externally-imposed perturbation in the homogenous
case) is described in polar coordinates (r, φ) by δΣ(r, φ),
which we decompose into Fourier modes according to
∞(cid:88)
δΣm(r)eim(φ−φp)(cid:105)
(cid:104)
δΣ(r, φ) =
Re
,
(5)
m=1
where φp = Ωpt is the azimuthal position of the planet
(in the homogeneous case, φp → 0). Each δΣm(r) is
a complex quantity describing the radial variation of the
amplitude and phase of the mode with azimuthal number
m. The radial velocity and azimuthal velocity perturba-
tions δur(r, φ) and δuφ(r, φ) are similarly expressed as
sums of Fourier modes δur,m(r) and δuφ,m(r).
3
is the squared Brunt -- Vaisala frequency, and DS = κ2 −
ω2 + N 2
r . Note that in the barotropic (uniform entropy)
limit, 1/LS → 0 and Nr → 0, equation (6) reduces to the
master equation of Goldreich & Tremaine (1979). The
surface density perturbation δΣm can be computed using
solutions of equation (6) according to
δΣm =
Σ
c2
s
δhm +
iΣ
LS ω
δur,m,
(9)
where the radial velocity perturbation δur,m is given in
terms of δhm and its radial derivative δh(cid:48)
m by
(cid:20)
δur,m =
−
i
DS
2mΩ
r
(cid:48)
ω(δh
m + Φ
(cid:48)
m)
(δhm + Φm) −
ω
LS
(cid:21)
δhm
.
(10)
The azimuthal velocity perturbation can be found in a
similar fashion,
(cid:21)
δhm
.
(cid:20) κ2
(cid:18)
2Ω
(cid:48)
m + Φ
(δh
(cid:48)
m)
(cid:19)
δuφ,m =
−
1
DS
mω
r
N 2
r
ω2
1 −
(δhm + Φm) −
κ2
2ΩLS
(11)
The components of the gravitational potential of the
planet are
Φm = −
GMp
rp
b(m)
1/2 (r/rp),
(12)
(13)
(cid:90) 2π
0
where bm
1/2 are (softened) Laplace coefficients,
b(m)
1/2 (α) =
1
π
cos(mψ)dψ
[1 − 2α cos(ψ) + α2 + 2]1/2 .
For the softening parameter, we choose = 0.6hp, a
value which is typically used in two-dimensional numer-
ical simulations of planet-disk interaction to mimic the
vertical averaging of the planetary gravity over the disk
In our calculations, as in Bae & Zhu (2018a),
height.
we ignore the "indirect" potential term, δm,1GMpr/r2
p,
which arises due to the motion of the central star around
the barycenter of the star + planet system. We motivate
this choice in Appendix A.
We solve equation (6) for a sufficient number of modes
to fully capture the two-dimensional structure of the sur-
face density perturbations. The mode solution method
closely follows that of Korycansky & Pollack (1993)
(KP93), as well as Rafikov & Petrovich (2012) and Petro-
vich & Rafikov (2012), and is described in detail in Ap-
pendix A. The solution is obtained on a logarithmic grid
with rin = 0.05rp and rout = 5.0rp, which, for our fidu-
cial parameters, has 6×104 grid points, or a resolution of
about 1300/H. All of the modes are solved on the same
grid to facilitate their synthesis, and the grid resolution
is set by the tight winding of the highest m modes near
the grid boundaries.
Examples of the surface density perturbation profiles
δΣm(r) for several low-order azimuthal modes are shown
in Fig. 1. Note that modes with higher m are more
tightly wound. Also note that here the m = 5 mode
has the largest amplitude. This is because the response
Fig. 1. -- Radial profiles of the fractional surface density per-
turbation for several low-order modes, for the case of the fiducial
parameters, hp = 0.1, q = 1, and p = 1.
(cid:26) d2
dr2 +
As a result of performing these steps, one arrives at the
linear equation describing the quantity4 δhm = δPm/Σ
(where δPm is the pressure perturbation) of the mode
with azimuthal number m (Baruteau & Masset 2008;
Tsang 2014; here we adopt the notation of the latter):
dr −
(δhm + Φm)
2mΩ
r ω
(cid:20) d
(cid:18) rΣ
dr
LSDS
ln
(cid:19)
(cid:18) ΣΩ
(cid:19)
DS
+
1
LS
(cid:21)
(cid:21)
δhm
+
2mΩ
LSr ω
(cid:19)(cid:21) d
(cid:19)(cid:27)
ln
(cid:20) d
(cid:18) rΣ
(cid:18) N 2
(cid:20) DS
dr
m2
r2
DS
r
ω2 − 1
+
c2
s
dΦm
1
LS
dr
= 0,
+
−
+
1
L2
S
+
1
LS
d
dr
ln
(6)
where ω = m(Ωp − Ω) is the Doppler-shifted frequency
of the tidal forcing due to the mth harmonic of the plan-
etary potential,
1
LS
=
1
γ
dS
dr
=
(γ − 1)p − q
γr
(7)
is the inverse length scale associated with the radial vari-
ation of entropy,
(cid:19)
dP
dr −
dΣ
dr
(cid:18) 1
c2
s
c2
s
LSr
N 2
r = −
1
Σ2
dP
dr
(q + p)
=
γ
(8)
4 For barotropic disks, δhm can be identified as the enthalpy
perturbation. However, in the presence of an entropy gradient, this
association no longer holds exactly. Instead, δhm simply serves as
a convenient variable for which a master equation can be found.
−0.4−0.20.00.20.4(a)m=2ReIm−0.4−0.20.00.20.4(Mp/Mth)−1δΣm/Σ(b)m=50.10.31.03.0r/rp−0.4−0.20.00.20.4(c)m=104
of the disk is dominated by modes with m close to
m∗ ≈ (2hp)−1 (here hp = 0.1 and m∗ = 5). The domi-
nant role of this characteristic m is related to the torque
cutoff phenomenon (Goldreich & Tremaine 1980). The
exact form of m∗ (i.e., the factor of 1/2) is not funda-
mental, but this choice is supported by our numerical
calculations.
Once
the mode
two-
dimensional surface density perturbation is then synthe-
sized according to
solutions are
found,
the
mmax(cid:88)
δΣm(r)eim(φ−φp)(cid:105)
(cid:104)
δΣ(r, φ) =
Re
,
(14)
m=1
where mmax is the value of m necessary to achieve a con-
verged perturbation structure; see Appendix A for de-
tails. The velocity perturbations δur(r, φ) and δuφ(r, φ)
are computed in the same manner. Note that the mode
perturbation, δΣm(r),
is a one-dimensional, complex
quantity (described by a radially-varying amplitude and
phase), while the synthesized perturbation, δΣ(r, φ) is a
two-dimensional, explicitly real quantity.
2.3. Parameters
The results of our calculations are fully determined by
four dimensionless parameters: hp, the disk aspect ra-
tio at the orbital radius of the planet, q, the power law
index of the disk temperature, p, the power law index
of the disk surface density, and the adiabatic index γ.
For the fiducial parameters, we choose hp = 0.1, q = 1
(corresponding to a constant disk aspect ratio h), p = 1,
and γ = 7/5. We find that our results are almost com-
pletely insensitive to the value of γ (see §3.3), and so
unless otherwise stated, we keep its value fixed. We
have explored the parameter space by performing calcu-
lations for which two of the three remaining parameters
are fixed at their fiducial values and the third is varied
over a plausible range of values for protoplanetary disks:
0.05 < hp < 0.15, 0 < q < 1, 0 < p < 3/2. We find that
the results are not sensitive to p, and so we primarily fo-
cus our analysis and discussion on the effects of varying
hp and q.
3. RESULTS FOR PLANET-DRIVEN SPIRALS
We will start presentation of our results with the case
of a spiral pattern driven by the gravity of an embed-
ded planet (forced or inhomogeneous case). To better
highlight new findings, we start by outlining the existing
picture of wave propagation in disks in §3.1. We then
describe general properties of the linear planet-driven
density waves found in our linear calculations (§3.2) and
provide a comparison with the results of direct numeri-
cal simulations (§3.4). These preliminaries form a basis
for subsequent in-depth discussion of the properties of
multiple spiral arms emerging in our calculations in §3.5.
3.1. Expectations Based on Simple Linear Theory
Existing linear calculations of the density wave prop-
agation in disks provide some guidance on the expected
outcome of our present calculation. Specializing to the
inhomogeneous case, the linear response of the disk is
expected to take a form of a one-armed spiral den-
sity wake (Rafikov 2002a; Ogilvie & Lubow 2002). For
r − rp (cid:29) Hp, the position of the wake is given approx-
imately by5
φlin = φp + sgn(r − rp)
Ω(r(cid:48)) − Ωp
cs(r(cid:48))
(cid:48)
,
dr
(15)
(cid:90) r
rp
and its amplitude (peak height) is given by, to within a
constant factor of order unity (Rafikov 2002a),
Mp
Mth
(cid:20) Ω(r) − Ωp
Ωp
(cid:21)1/2(cid:20) cs(r)
(cid:21)−3/2
cs(rp)
Σp
Σ(r)
rp
r
δΣlin
Σ(r)
=
×
Here
Mth = h3
pM∗
(16)
.
(17)
is the thermal mass. The first (constant) factor on the
right-hand side of Equation (16) describes the initial am-
plitude of the density wake formed within a few scale
heights of the planet (in the homogeneous case consid-
ered in §4, rp is replaced with the radius at which the
perturbation is imposed, and Mp/Mth with an arbitrary
constant), while the other factors indicate how the ampli-
tude varies as the wake propagates away from the planet.
The radial scaling is dictated by the conservation of an-
gular momentum flux (AMF),
(cid:73)
FJ (r) = r2Σ(r)
δur(r, φ)δuφ(r, φ)dφ,
(18)
which in the absence of explicit dissipation (linear or
nonlinear) must be constant far from the planet, outside
the wave excitation region (Goodman & Rafikov 2001;
Rafikov 2002a). The characteristic scale of FJ , result-
ing from the sum of the one-sided Lindblad torques, is
(Goldreich & Tremaine 1980; Ward 1997)
FJ,0 =
−3
p Σpr4
h
pΩ2
p.
(19)
(cid:18) Mp
(cid:19)2
M∗
These known linear results will be used as a reference for
comparison in our current calculations.
3.2. Results of Linear Calculations: General Properties
In Figure 2 we show two-dimensional maps of the per-
turbed surface density (in polar coordinates r, φ) result-
ing from our typical linear calculations. The case with
the fiducial parameters is shown (Fig. 2a), as well as sev-
eral other cases, discussed in §3.6, in which each of the
three parameters (indicated at the top of the panel) is
varied from its fiducial value.
The same maps of the perturbed surface density are
also shown in Cartesian coordinates in Figure 3, in or-
der to demonstrate the true geometry of the spirals as
seen by an observer sensitive to the gas surface density
perturbations. Note that in Figure 3, only the part of
the disk interior to the orbit of the planet is shown, and
the range of color scale for the perturbations is smaller
5 Note that analogous expression in Rafikov (2002a) has a dif-
ferent sign of the second term.
5
Fig. 2. -- The two-dimensional fractional surface density perturbation (scaled by the ratio of the planet mass to the thermal mass), shown
in polar coordinates, for the case of the fiducial parameters (hp = 0.1, q = 1, p = 1; left panel), and for several other cases, demonstrating
the effect of varying each parameter: the disk aspect ratio, hp (middle-left panel), the temperature power-law index, q (middle-right panel),
and the surface density power-law index, p (right panel). The positions of the primary arm, secondary arm, and tertiary arm are indicated
in the leftmost panel.
than in Figure 2, in order to focus on the detail of the
perturbations in the inner disk.6
Also, to highlight the details of the spiral arm evolu-
tion, in Figure 4a,b we display azimuthal profiles of δΣ
at a fixed radius r, normalized by the linear wake ampli-
tude δΣlin given by equation (16). These profiles can be
thought of as horizontal cuts in Figure 2.
Generically, in the outer disk (r > rp), a single strong
spiral arm, i.e., a narrow structure with a δΣ > 0 peak
is present.
Its azimuthal profile remains relatively un-
changed as the arm winds up with increasing r, see Fig-
ure 4a. It is accompanied by a comparatively wider sur-
face density trough, with δΣ < 0, which trails behind the
arm (in the sense of the Keplerian rotation of the disk).
The case with a smaller disk aspect ratio, hp = 0.05,
represents an exception, as it develops a weak additional
spiral arm beyond r ≈ 3rp, although this is not clearly
discernible in Figure 2 (see §3.5 for details).
In the inner disk (r < rp), there is a single spiral arm
(the "primary spiral") near the planet, accompanied by
a trough which leads the arm. However, farther from
6 We point out that radially narrow features seen at r ≈ rp in
Figs. 2 -- 3 are perturbations in the corotation region which result
from the presence of a radial entropy gradient (i.e., due to terms
depending on 1/LS in equation 6). As they are very localized
near the orbit of the planet and have no effect on the global spiral
structure of the disk, we do not examine these features in detail in
this work.
the planet, the structure of the density wake deviates
from the simple behavior seen in the outer disk. For the
fiducial parameters, the presence of second peak with
δΣ > 0 (the "secondary spiral arm") becomes apparent
at r ≈ 0.5rp, accompanied by a deepening of the ini-
tial trough, see Figure 4b. Towards smaller radii, the
strength of the secondary spiral increases, and it also be-
comes accompanied by a leading trough. At even smaller
radii (r (cid:46) 0.1rp), a third arm (the "tertiary spiral")
forms. These multiple spirals are robustly present for a
variety of disk parameters. Note that the additional spi-
rals (secondary, tertiary, and so on) are always located
ahead of the primary spiral.
In Figure 4c,d we also show the azimuthal distribution
of angular momentum flux,
fJ (r, φ) = r2Σ(r)δur(r, φ)δuφ(r, φ),
(20)
at different radii. This quantity is related to the angular
momentum flux according to
(cid:73)
FJ (r) =
fJ (r, φ)dφ.
(21)
The evolution of the azimuthal profile of fJ with r al-
lows us to trace the exchange of angular momentum flux
between the multiple arms as the wake propagates away
from the planet.
The AMF is peaked near the locations of both the
−0.50.00.5(φ−φp)/π0.10.31.03.0r/rp(a)FiducialPrimaryPrimarySecondaryTertiary−0.50.00.5(φ−φp)/π(b)hp=0.05−0.50.00.5(φ−φp)/π(c)q=1/2−0.50.00.5(φ−φp)/π(d)p=0−2.0−1.5−1.0−0.50.00.51.01.52.0(Mp/Mth)−1δΣ/Σ6
Fig. 3. -- The two-dimensional fractional surface density perturbation for the fiducial paramaters and several cases with varied parameters,
as in Figure 2, but shown here in Cartesian coordinates, and focusing only on the spiral structure in the inner disk, interior to the orbit of
the planet. The positions of the primary, secondary, and tertiary arms are indicated in the top-left panel.
peaks (δΣ > 0) and the troughs (δΣ < 0) of the sur-
face density perturbation, since the angular momentum
flux is approximately proportional to the square of δΣ
far from the planet (Rafikov 2002a),
rc3
s
(δΣ)2.
fJ ≈
(22)
In particular, in the outer disk, for r/rp (cid:38) a few, the
trough trailing the primary arm of the surface density
perturbation carries a significant fraction of the AMF.
Ω − ΩpΣ
We provide more in depth discussion of the different
features of the density wake in §3.5.
3.3. Dependence on Adiabatic Index
In the calculations presented in this section, we have
adopted γ = 7/5 as the fiducial adiabatic index. We
have also carried out calculations for hp = 0.1, q = 1,
and p = 1 with different values of γ, the results of which
are shown in Figure 5. Here azimuthal profiles of δΣ are
shown at different radii for several values of γ in addi-
tion to γ = 7/5. The case γ = 1.001 corresponds to an
almost (locally) isothermal disk. The case γ = 4/3 repre-
sents the effective two-dimensional adiabatic index corre-
sponding to a three-dimensional adiabatic index Γ = 7/5.
These different adiabatic indices are related according to
γ = (3Γ − 1)/(Γ + 1) (Goldreich et al. 1986; Ostriker
et al. 1992) in the low-frequency limit, ω (cid:28) Ω (valid for
r ≈ rp). For γ = 2, the disk has a uniform entropy profile
(see equation 7) and we recover the barotropic limit.
The surface density profiles shown in Figure 5 demon-
strate that our results are remarkably insensitive to the
value of γ7. Mathematically, this is because γ-dependent
contributions appear in equation (6) only through terms
varying on scales of ∼ rp. In §5, we show that the forma-
tion of multiple spirals is well-described by considering
7 The perturbation structure in the immediate vicinity of rp
(within about 0.1H) has a strong dependence on γ, but this is not
important for our present study of spiral arms far from the planet.
−0.50.00.5y/rp(a)FiducialPrimarySecondaryTertiary(b)hp=0.05−0.50.00.5x/rp−0.50.00.5y/rp(c)q=1/2−0.50.00.5x/rp(d)p=0−1.0−0.50.00.51.0(Mp/Mth)−1δΣ/Σ7
Fig. 4. -- Top panels: profiles of the surface density perturbation δΣ, scaled by δΣlin (see equation 16) to facilitate comparison between the
profiles at different radii. Bottom panels: profiles of the azimuthal distribution of angular momentum flux fJ , in terms of the characteristic
angular momentum flux FJ,0 (equation 19). Profiles are shown for several different radii in the outer disk (left panels) and the inner disk
(right panels), for the fiducial disk parameters.
the phases of different modes in the local (WKB) limit,
in which the globally-varying terms in equation (6) with
explicit dependence on γ are unimportant. Therefore, it
is sufficient to only consider the fiducial γ = 7/5 in our
subsequent discussion.
3.4. Numerical Validation
To validate and understand the limitations of our semi-
analytical linear calculations, we have also carried out a
set of direct numerical simulations of planet-disk inter-
action in the low mass regime using fargo3d (Ben´ıtez-
Llambay & Masset 2016). These simulations will be dis-
cussed in detail in a future work. Here we describe only
the basic setup as well as the results for the fiducial disk
parameters, hp = 0.1, q = 1, and p = 1.
We choose a planet mass Mp = 10−5M∗, or 0.01Mth.
Since Mp (cid:28) Mth, the linear regime is appropriate for
describing the disk response, except far ((cid:38) 6Hp) from
the planet, where nonlinear effects (fully captured in
simulations) -- wake evolution into a shock and subse-
quent dissipation -- become non-negligible (Goodman &
Rafikov 2001). A softening length of 0.6Hp is applied to
the potential of the planet. We use a logarithmically-
spaced radial grid extending from rin = 0.05rp to rout =
5.0rp with Nr = 4505 grid cells, and a uniformly-spaced
azimuthal grid with Nφ = 6144 grid cells. The resulting
grid cells have a roughly square shape with a resolution
of 98 cells per scale height. Wave damping zones (e.g.,
de Val-Borro et al. 2006) are implemented for r < 0.06rp
and r > 4.5rp in order to minimize wave reflection at the
boundaries (note that direct comparison with the results
of our linear calculations is not possible in the damping
zones).
We adopt an ideal equation of state with γ = 1.001.
This value of γ close to unity is chosen in order to avoid
any significant heating of the disk due to wave dissipation
during our simulations. Note that although this setup re-
sembles that of a locally isothermal disk, the two cases
are physically distinct. In locally isothermal disks, non-
axisymmetric perturbations are known to exchange an-
gular momentum with the background flow (Lin & Pa-
paloizou 2011; Lin 2015), which can lead to anomalous
results.8 No explicit viscosity is included in the simula-
tion.
We compare the results of the numerical simulation
after 10 orbits (when steady state is reached) with our
linear calculations (using γ = 1.001 in order to facilitate
direct comparison) in Figure 6. The profiles of the sur-
face density perturbation δΣ are shown at several differ-
ent disk radii in the inner disk, highlighting the multiple
spiral arm structure. The profiles are rotated by φlin(r)
and scaled by δΣlin(r), as in Figure 4.
The numerical results exhibit very good agreement
with the linear prediction. The positions of the primary,
secondary, and tertiary arms are closely reproduced, dif-
fering by (cid:46) 0.05 radians from the linear prediction even
at the smallest radii. In particular, the initial secondary-
8 We look into this issue in more detail in Miranda & Rafikov
(in prep.).
−1.0−0.50.00.51.01.52.02.5δΣ/δΣlinOuterDisk(a)r=1.10rpr=1.50rpr=3.00rpr=5.00rp−1.0−0.50.00.51.01.52.02.5InnerDisk(b)r=0.90rpr=0.50rpr=0.25rpr=0.10rp−1.0−0.50.00.51.0[φ−φlin(r)]/π−1.00.01.02.03.04.05.06.0fJ/FJ,0(c)−1.0−0.50.00.51.0[φ−φlin(r)]/π−0.20.00.20.40.60.81.01.2(d)8
Fig. 5. -- Surface density perturbation profiles in the inner disk
for the fiducial disk parameters, hp = 0.1, q = p = 1, and different
values of the adiabatic index γ.
to-primary arm separation is ≈ 60◦, and this separation
decreases towards smaller radii, in agreement with our
calculations (see §3.5). This is a unique feature of the
secondary spiral in the linear regime, which should be
contrasted with the 180◦ separation which arises when
Mp > Mth (e.g., Zhu et al. 2015). The amplitudes of the
arms and troughs in the numerical simulation also show
good agreement with the linear prediction, exhibiting es-
sentially negligible differences, except for at small radii
(r (cid:46) 0.1rp), where differences of ≈ 20% arise due to non-
linear effects (especially in the amplitude of the trough
between the primary and secondary spirals). Notably,
the amplitude of the secondary arm overtaking that of
the primary for r (cid:46) 0.1rp is well reproduced in the simu-
lation. The agreement between the numerical simulation
and linear theory supports full applicability of our results
to low-mass planets.
We note that no shocks are present in Figure 6, in
contradiction with the results of Rafikov (2002a), which
predict shock formation at about 0.3rp. However, the
calculations of Rafikov (2002a) were carried out for the
case of a single spiral arm propagating in a self-similar
fashion. In our calculations, there is an exchange of an-
gular momentum flux from the primary spiral arm to the
secondary spiral arm, which modifies this picture. Evi-
dently this exchange lowers the amplitude of the primary
Fig. 6. -- Profiles of the surface density perturbation at several
radii in the inner disk, comparing the results of our numerical
simulation (solid curves) and the results of our linear calculation
(dashed curves) for the fiducial parameters (with γ = 1.001).
arm and suppresses shock formation for the planet mass
(Mp = 0.01Mth) that we have considered here.
3.5. Structure of the Multiple Spirals
We now characterize the morphological properties --
amplitude, width, and arm-to-arm separation -- of the
spiral arms, the radial dependence of these properties,
and how they vary with the disk parameters. We first
briefly describe the procedure for identifying and char-
acterizing the spiral arms, the results of which are illus-
trated in Figures 7 -- 8. We emphasize that the procedure
for decomposing δΣ into discrete spiral arms and troughs
is heuristic and not unique; in practice this decomposi-
tion can be done using different criteria than the ones
we used. Therefore the results should be taken as semi-
quantitative descriptions.
3.5.1. Identification and Characterization of Spiral Arms
At r = rp, δΣ has a single (global) maximum, located
at ≈ φp, which we identify as the primary arm. By
following the position of this maximum across different
radii, we obtain the position of the primary arm as a
function of r. If at some r, a second (local) maximum
with an amplitude equal to at least 10% of the ampli-
tude of the primary arm is present, it is identified as the
secondary spiral arm. This relative amplitude threshold
is necessary in order to avoid spurious "detections" of
additional spirals. The position of the secondary is then
0.00.51.0(a)r=0.80rpγ=1.001γ=4/3γ=7/5γ=2−0.50.00.51.0(b)r=0.40rp−0.50.00.5δΣ/δΣlin(c)r=0.20rp−0.50.00.5(d)r=0.10rp−1.0−0.50.00.51.0[φ−φlin(r)]/π−0.50.00.5(e)r=0.05rp0.00.51.0(a)r=0.80rpTheorySimulationTheorySimulation−0.50.00.51.0(b)r=0.40rp−0.50.00.5δΣ/δΣlin(c)r=0.20rp−0.50.00.5(d)r=0.10rp−1.0−0.50.00.51.0[φ−φlin(r)]/π−0.50.00.5(e)r=0.06rp9
Fig. 7. -- Properties of the multiple spiral arms (solid curves) and their associated troughs (dashed curves) as functions of radius, for
different values of the disk aspect ratio near the planet, hp (left, right, and middle panels), and with fixed values of the temperature and
surface density power law indices q = 1 and p = 1. Top panels: Positions of the arms and troughs relative to the primary arm. Top
middle panels: The full-width half maxima of the arms and troughs. Bottom middle panels: The absolute values of the surface density
perturbations associated with the arms and troughs, scaled by δΣlin (equation 16). Bottom panels: The angular momentum flux FJ
associated with each arm or trough, as well as the sum of the angular momentum flux of each arm and its associated trough (dotted
curves), and the total angular momentum flux (sum over all of the arms and troughs; dot-dashed curves).
followed across different radii in the same manner as the
primary. The emergence of subsequent (e.g., tertiary,
quaternary) arms is determined by a similar criterion,
except we require the amplitude to be at least 10% of
the strongest spiral arm at a given r, which may not nec-
essarily be the primary (since it amplitude decays with
distance from the planet). The positions of the spiral
arms are denoted φP, φS, and so on.
We also identify the troughs, i.e., local minima of δΣ.
We associate each trough with a spiral arm at each r,
since features tend to appear in arm/trough pairs (see
Figure 4). The troughs are therefore designated as the
"primary trough", "secondary trough", and so on. Note
that near the planet, the primary arm is accompanied by
two troughs, one leading and one trailing. Toward the
outer disk, the trailing trough becomes the stronger of
the two, and so it is identified as the primary trough. In
the inner disk, the leading trough becomes stronger, and
so it is identified as the primary trough instead. This
is the origin of the discontinuity in the position of the
primary trough in Figs. 7a -- c and 8a -- b. Multiple arms
formed in the inner disk are each similarly accompanied
by a leading trough.
In the outer disk, when multiple
arms are present (we find at most two, and only for our
thinnest disk), they are instead accompanied by trailing
troughs.
Using the profiles of δΣ in the vicinity of the peaks and
troughs, we quantify the amplitudes and widths of the
arms/troughs using the maximum and the full width at
half maximum (FWHM) of δΣ for arms, and of −δΣ for
troughs. The arm amplitudes are denoted by δΣP, δΣS,
etc.
Note that since troughs are simply identified as local
minima, it is sometimes the case (for high order troughs,
e.g., tertiary and beyond) that δΣ > 0 at the trough
location. In this case the width of the trough is unde-
fined, since the amplitude is measured relative to zero.
But as the trough evolves, eventually δΣ at the mini-
mum becomes negative, so the trough has a well-defined
(but narrow) width. The trough width increases as its
amplitude grows, until it resembles the initial widths of
lower-order troughs, and then becomes more narrow to-
wards the inner disk, in accordance with the behavior
of the other arms/troughs. This explains the anomalous
behavior of some of the high-order trough widths seen
in, e.g., Figure 7f and 8c -- d.
We also keep track the AMF FJ carried by each arm
or trough by integrating the AMF distribution (equa-
tion 20) over its azimuthal extent, which is delineated by
the zeros of δΣ. These roughly correspond to the zeros
of fJ , since the latter is approximately proportional to
(δΣ)2 (see Fig. 4 and equation 22). Note that the AMF
090180270360φ−φP[◦](a)hp=0.05020406080FWHM[◦](d)0.01.02.03.0δΣ/δΣlin(g)0.10.30.51.03.0r/rp0.00.20.40.60.8FJ/FJ,0(j)(b)hp=0.1(e)(h)0.10.30.51.03.0r/rp(k)(c)hp=0.15Primary(Arm)Primary(Trough)SecondaryTertiaryQuaternary(f)(i)0.10.30.51.03.0r/rp(l)Arm+TroughTotal10
Fig. 8. -- The same as Figure 7, but for different values of the temperature power law index q, with hp = 0.1.
Fig. 9. -- The radii at which the amplitude of the surface density perturbation associated with the secondary spiral arm, δΣS, relative to
the amplitude of the primary arm, δΣP, is equal to 0.2 (red points), 0.5 (blue points), and 1 (green points). In each panel, two of three disk
parameters (aspect ratio hp, temperature power law index q, and surface density power law index p) are fixed at their fiducial values, while
the value of the third parameter is varied. The dashed curves indicate the theoretical predictions given by equations (38) -- (39), calibrated
to the fiducial disk parameters (i.e., using hp = 0.1).
is divided up and assigned to the arms and troughs in a
conservative manner, so that the sum of the AMF asso-
ciated with all of the peaks and troughs at a given radius
is equal to FJ (r).
The extracted properties of the arms and troughs (po-
sition, width, amplitude, and AMF) are shown for the
fiducial case and for cases with different disk aspect ra-
tios in Figure 7, and for cases with different temper-
ature power law indices in Figure 8. Note that the
arm/trough positions are shown relative to the position
of the primary arm, φP, which is well-approximated by
equation (15) (although it may differ by ∼ 10◦ far from
the planet, see Fig. 4). Also note that the amplitudes are
scaled by δΣlin, in order to remove the simple variation
of amplitude with r resulting from angular momentum
flux conservation. The AMF (lower panels) is scaled by
the characteristic AMF due to the one-sided Lindblad
torques given by Equation (19).
3.5.2. Outer Disk
The structure of the surface density perturbation (see
Figs. 7g -- i and 8e -- f) in the outer disk is fairly sim-
ple. Typically the perturbation consists only of a pri-
mary arm, with a width of about 10◦, and its associ-
090180270360φ−φP[◦](a)q=0020406080FWHM[◦](c)0.00.51.01.52.0δΣ/δΣlin(e)0.10.30.51.03.0r/rp0.00.20.40.6FJ/FJ,0(g)(b)q=1/2Primary(Arm)Primary(Trough)SecondaryTertiaryQuaternaryQuinary(d)(f)0.10.30.51.03.0r/rp(h)Arm+TroughTotal0.00.20.40.60.050.100.15r/rphp(a)q=1,p=1δΣS/δΣP=0.2δΣS/δΣP=0.5δΣS/δΣP=10.00.51.0q(b)hp=0.1,p=10.00.51.01.5p(c)hp=0.1,q=1TheoryNumerical11
Fig. 10. -- Two-dimensional fractional surface density perturbations for the propagation of passive spiral wake with an initially specified
profile at rout (see §4). The leftmost panel shows the case of the fiducial disk parameters hout = 0.1, q = 1, and p = 1, and σ = 0.06. In the
other panels, the disk aspect ratio, temperature power law index, and initial width are varied. In all cases, the single spiral arm imposed
at rout splits up into two or more arms as it propagates inwards.
ated trough, which is about four times wider. Their
amplitudes vary slowly, with the gradual decay of the
arm amplitude accompanied by the gradual growth of
the trough amplitude. The amplitude of the trough re-
mains small relative to the arm (a few tenths at most),
although it carries a significant fraction of the AMF (see
Figs. 7j -- l and 8g -- h) due to its large width. For hp = 0.1
and smaller (with q = 1), the trough carries most of the
AMF (i.e., more than the peak) for r (cid:38) (1.5 − 2.0)rp.
For the thinnest disk we have considered (hp = 0.05),
a secondary arm does form at about 3rp (see Fig. 7g),
which is explained in §5.2.2. Its amplitude grows very
slowly with r, and is only about 12% of that of the pri-
mary at 5rp. Note that we have solved for the structure
of the perturbations only out to 5rp, and so it is pos-
sible that a secondary arm emerges beyond this radius
for larger hp as well (although our analytic estimates in-
dicate that this is unlikely, see §5.2.2).
If so, it forms
very far from the planet in comparison to the inner disk,
where a secondary arm always forms at about half of the
orbital radius of the planet. And as demonstrated by the
hp = 0.05 case, the outer secondary spiral is very weak
even when it exists. Therefore we do not devote much
attention to secondary spirals in the outer disk.
3.5.3. Inner Disk
In the inner disk, multiple (three to five) spiral arms
are robustly formed for a variety of parameters. For the
fiducial parameters (Figure 7b,e,h,k), the secondary arm
first appears at 0.46rp, with an initial separation of about
70◦ from the primary arm, and an initial width of 45◦.
At smaller radii, its separation relative to the primary
arm decreases to about 30◦, and its width decreases to
about 20◦. The amplitude of the secondary arm increases
as that of the primary decreases. At r = 0.17rp, the
amplitude of the secondary is about half of the primary,
and at r = 0.06rp, the secondary amplitude exceeds the
primary amplitude.
A weak tertiary arm forms at about 0.3rp, and a weak
quaternary arm forms at about 0.1rp. The tertiary and
quaternary arms each form at ≈ 135◦ from the previ-
ous arm (roughly twice the secondary-to-primary sepa-
ration), with an initial width of about 60◦, see Figure
7b,e. After the quaternary arm forms, roughly the entire
2π extent of the disk is populated with spirals, and so it
is unlikely that any more well-defined arms would form9.
Nonetheless, we emphasize that the tertiary and higher-
order arms have very small amplitudes relative to the
primary and secondary even at r = 0.05rp (see Figure
7h), and thus are relatively unimportant features.
As in the outer disk, the primary trough carries a sig-
nificant portion of the total AMF (see Figs. 7j -- l and 8g --
h), and typically carries more AMF than the primary
arm at r (cid:46) (0.5−0.8)rp, which is a larger radius than the
9 Although since the arms become more narrow toward the in-
ner disk, there may possibly be room for additional arms at even
smaller radii.
−0.50.00.5φ/π0.10.30.51.0r/rout(a)Fiducial−0.50.00.5φ/π(b)hout=0.05−0.50.00.5φ/π(c)q=1/2−0.50.00.5φ/π(d)σ=0.18−1.0−0.50.00.51.0δΣ/Σ12
panel, two of the three disk parameters (hp, q, and p) are
held fixed while the third one is varied. These properties
depend most strongly on hp, with the arm forming closer
to planet for smaller hp. For values of q smaller than the
fiducial q = 1, i.e., for more flared disks, the secondary
arm forms somewhat closer to the planet. However, its
amplitude grows with radius faster than the case with
q = 1, reaching 50% and 100% of the primary amplitude
at significantly larger radii.
The development of the secondary spiral is relatively
insensitive to the slope of the surface density profile, for
a plausible range of values (0 < p < 3/2). This is evident
in the structure of the 2D surface density maps in Fig-
ures 2 -- 3, which show the cases p = 1 and p = 0 (note that
the overall scaling of the amplitude of the perturbations
with r differs between the two cases, but this is triv-
ially described by AMF conservation, see equation 16).
This is shown quantitatively in Figure 9c. The radius at
which the secondary spiral forms differs by only about
5% between the cases p = 0 and p = 3/2.
Fig. 11. -- Profiles of the surface density perturbation (top panel)
and azimuthal distribution of angular momentum flux (bottom
panel) at different radii, as in Figure 4, but for the homogeneous
problem of a passive spiral wake.
one at which the secondary arm forms, or at which the
secondary arm amplitude becomes stronger than the pri-
mary (in terms of amplitude). The AMF of the secondary
arm exceeds primary arm AMF for r (cid:46) (0.4 − 0.5)rp.
Far inside inner disk (r ≈ 0.1rp), the AMF is primarily
carried by secondary arm, secondary trough, and pri-
mary trough, while contributions from the other peaks
and troughs are negligible.
3.6. Dependence on Disk Parameters
Figure 2a,c shows that lowering hp results in more
tightly wound spiral arms. For such colder and thinner
disks, the secondary arm (in the inner disk) forms closer
to the primary (in azimuthal angle), with a smaller ini-
tial width and smaller initial arm-to-arm separation, and
it becomes stronger than the primary also at a larger ra-
dius, see Figure 7a,d,g,j.
For temperature profiles with q different from unity,
Figure 2a,b shows that the pitch angle of the spirals
vary with radius. For q < 1, the initial properties of
the secondary arm are relatively unchanged, but its am-
plitude becomes larger than the primary amplitude at
larger radii (similar to the case of lower hp), see Fig-
ure 8. Additional (tertiary and beyond) arms also form
closer to the planet for thinner disks, and more arms are
formed -- as many as five for hp = 0.05 (however, the full
azimuthal extent of the inner disk is not fully populated
with spirals in this case, as it is for larger aspect ratios).
The development of the secondary arm in the inner
disk is summarized in Figure 9 for different disk parame-
ters. It shows the radial location at which the amplitude
of the secondary arm, δΣS, relative to the amplitude of
the primary, δΣP, is equal to 0.2, 0.5, and 1.
In each
4. RESULTS FOR PASSIVELY PROPAGATING SPIRAL
WAVES
As will be shown in §5, the formation of the secondary
spiral can be described fairly well by the local (WKB)
approximation, which is valid far from resonances, where
multiple spiral arms emerge, and where the gravitational
influence of the planet is negligible. This indicates that
the external potential of the planet plays a minimal role
in this process. Therefore, motivated by the results of
Arzamasskiy & Rafikov (2018), we may expect that any
spiral wake formed by the phase coherence of a series of
modes launched in a companion-free disk will also break
up into multiple spirals.
To this end, we investigate the passive linear propaga-
tion of a spiral wake (not subject to the external potential
of a planet). More specifically, we perform a calculation
representing a linear analogue of the nonlinear calcula-
tion of Arzamasskiy & Rafikov (2018), closely following
their setup. In this setup, a surface density perturbation
δΣ(φ) rotating with a pattern frequency ωp is imposed
at a radius rout, and is allowed to propagate towards the
disk center (only inward propagation is allowed provided
that ωp < Ωout).
We solve for the structure of the modes of the homo-
geneous wave equation (6) with Φm = 0 by shooting so-
lutions inward from rout in order to satisfy the outgoing
wave boundary condition at rin = 0.05rout. For simplic-
ity, here we choose a mode pattern frequency ωp = 0
(so that the wave pattern is fixed in the inertial frame),
which corresponds to waves launched at a radius (cid:29) rout.
Once the mode profiles are obtained, any azimuthal pro-
file δΣ can be constructed from an appropriate linear
combination of the different δΣm's.
The perturbation imposed at rout is chosen to be a
−0.50.00.51.0δΣ/δΣlin(a)r/rout=1.00r/rout=0.50r/rout=0.25r/rout=0.10−1.0−0.50.00.51.0[φ−φlin(r)]/π0.02.04.06.08.010.0fJ/FJ,out(b)13
Fig. 12. -- The radii at which the amplitude of the secondary arm reaches 0.2 (red points), 0.5 (blue points), and 1 (green points) relative
to that of the primary arm, as in Figure 9, but for the homogeneous problem. In the different panels, the aspect ratio, temperature power
law index, and initial perturbation width are varied while the other parameter values are held fixed. Note that "missing" points indicate
that the secondary amplitude did not reach the specified threshold anywhere for r > rin = 0.05 for the specified parameter value. The
dashed curves show the theoretical predictions given by equations (38) and (44), calibrated to the fiducial parameters.
Gaussian10,
(cid:34)
δΣ(rout) ∝ exp
−
(cid:19)2(cid:35)
.
(cid:18) φ
σ
1
2
(23)
Thus, there are four parameters in our setup. These
are the three disk parameters hout and q which define
the sound speed profile, now parameterized by cs(r) =
houtroutΩout(r/rout)−q/2, and the surface density power
law index p, as well as the initial azimuthal width σ of
the surface density perturbation. For the fiducial param-
eters, we choose hout = 0.1, q = 1, p = 1, and σ = 0.06.
The fiducial σ is chosen to produce a surface density
profile similar to the one produced near the planet for
the case of the fiducial parameters in the inhomogeneous
problem.
Figure 10 shows 2D maps of the fractional surface den-
sity perturbation, for the fiducial parameters, and for
several cases with varied parameters. In all of these cases,
the initial, single peaked perturbation profile breaks up
into multiple arms, in agreement with findings of Arza-
masskiy & Rafikov (2018). The azimuthal profiles of δΣ
at different radii are shown in Figure 11 for the fiducial
parameters. They are qualitatively very similar to the
profiles that develop in the inner disk in the case of the
spiral wake driven by a planet, see Figure 4. The az-
imuthal distribution of the angular momentum flux sim-
ilarly becomes divided up into several different peaks.
We emphasize that the appearance of multiple spirals in
this setup, in which there are no resonances in the disk,
indicates that the formation of multiple spirals is not
connected to resonances.
Figure 12 summarizes the dependence of secondary
arm amplitude on disk parameters and the initial width
10 The actual profile imposed at rout differs slightly from the
form specified by equation (23). Specifically, the actual profile
is "missing" the m = 1 Fourier component. This is because the
m = 1 mode is evanescent interior to its OLR (which is assumed to
be located at r (cid:29) rout), and exterior to its ILR (which is formally
located at r = 0). Therefore, the amplitude of the m = 1 mode is
assumed to be zero at rout, as a result of its evanescent propagation
from large r. Note that if we had instead chosen the amplitude of
the m = 1 mode to be finite at rout (so that δΣ is given exactly by
eq. 23), it would decay away towards small r anyway, and it would
not strongly affect our results.
of the spiral wake. The dependence on the surface den-
sity profile index p is extremely weak. In fact, it is much
weaker than in the case with a planet: the radius at
which the secondary spiral emerges changes by less than
1% when the value of p is varied between 0 and 1. This
lack of sensitivity to Σ profile is explained in §5.
On the other hand, properties of the secondary spiral
depend strongly on the initial width of the perturbation,
see Figure 12c. Wider initial perturbations lead to the
development of the secondary spiral at larger radii, which
is discussed in §5. In this experiment, we have isolated
the effect of the initial perturbation width, as it is con-
trolled by the parameter σ, which is unrelated to the disk
parameters. In the case of a density wave launched by
a planet, the width of the spiral wake near the planet
is set by the disk aspect ratio, since the perturbation
is dominated by modes with m ∼ 1/hp, and so thinner
disks produce narrower initial wakes.
In the homoge-
neous case, by separating these two effects, we see that
variation of the disk aspect ratio has an even stronger
effect on the emergence of the secondary spiral than it
does in the case with a planet: changing hout by a fac-
tor of three changes the radius at which the secondary
emerges by a factor of two. Evidently these two effects
somewhat cancel out in the case with a planet.
5. ANALYTICAL UNDERSTANDING OF OUR RESULTS
In this section, we present theoretical arguments, fol-
lowing those made by Ogilvie & Lubow (2002) and Bae
& Zhu (2018a), which interpret spiral arm formation as
resulting from constructive interference among different
azimuthal modes. We then use these arguments to ex-
plain and interpret the results of our numerical calcula-
tions.
5.1. Mode Phases and Interference
The surface density perturbation of the mode with az-
imuthal number m is
δΣm(r, φ) = δΣm(r) exp[iψm(r)],
where
ψm(r) = Arg[δΣm(r)] + m(φ − φp).
(24)
(25)
0.00.20.40.60.050.100.15r/routhout(a)q=1,σ=0.060.00.51.0q(b)hout=0.1,σ=0.06δΣS/δΣP=0.2δΣS/δΣP=0.5δΣS/δΣP=10.050.100.150.20σ(c)hout=0.1,q=1TheoryNumerical14
The phase of δΣm can be expressed as (Ogilvie & Lubow
2002)
Arg[δΣm(r)] = sgn(r − rp)
where
r± =
(cid:90) r
(cid:19)2/3
+
r±
π
4
(cid:18)
1
m
1 ±
(cid:48)
(cid:48)
,
)dr
km(r
(26)
rp
(27)
is the location11 of the outer (+)/inner (−) Lindblad
resonance, and the π/4 term is the phase shift associated
with the resonance. The radial number is given in the
local (WKB) limit (and in the Keplerian limit κ = Ω) by
(cid:2)m2(Ω − Ωp)2 − Ω2(cid:3)1/2
km(r) =
1
cs
Note that for m (cid:29) 1, km(r) ≈ km(r), where
km(r) =
m
cs Ω − Ωp,
which is simply proportional to m. Defining
we can write
∆km(r) = km(r) − km(r),
(cid:90) r±
(cid:90) r
+ m(φ − φP)
(cid:48)
km(r
ψm = sgn(r − rp)
(cid:48)
∆km(r
)dr
π
4
+
−
rp
r±
.
(28)
(29)
(30)
(cid:48)
(cid:48)
.
)dr
(31)
Here we have defined12
φP = φp −
= φp −
(cid:90) r
(cid:90) r
rp
rp
(cid:48)
(cid:48)
)dr
km(r
Ω(r(cid:48)) − Ωp
cs(r(cid:48))
(cid:48)
.
dr
(32)
Note the distinction between φp (with a lowercase sub-
script), the position of the planet, and φP (with an upper-
case subscript), the position of the primary spiral arm.
Also, in equations (26), (31), (32) note the different lim-
its of integration.
The perturbation δΣm has a maximum found by set-
ting ψm = 0. For modes with m (cid:29) 1, the last two terms
in equation (31) become negligible, so that the maxima
of these modes have positions φ ≈ φP. For general m,
the maximum of δΣm interferes constructively with these
modes if (cid:90) r
∆km(r
(cid:48)
(cid:48)
)dr
−
r±
(cid:48)
km(r
) < ∆φ0,
(33)
where ∆φ0 is the maximum phase difference which re-
sults in constructive interference. In general, this crite-
rion is indeed satisfied (for some range of r and values of
m), resulting in the formation of the primary spiral arm.
Therefore, φP gives the approximate position of the pri-
mary spiral (Ogilvie & Lubow 2002; Rafikov 2002a). To
11 Note that this expression neglects O(h2
p) terms due to radial
pressure support and non-zero Nr.
12 Note that φP is equivalent to φlin as defined in equation (15).
Here it has been redefined to emphasize that it gives a theoretical
prediction for the azimuthal position of the primary spiral arm.
(cid:90) r±
rp
understand the origin of higher-order arms (secondary,
tertiary, etc.), let us note that maxima of δΣm are also
attained at ψm = 2πn, where n = 1, 2, . . . , m − 2, m − 1
is an integer. Maxima in this range of n are distinct in
a sense that, at a given r, each of them corresponds to
a well defined azimuthal location φ = φm,n (in a frame
co-rotating with the planet), where
φm,n(r) = φp +
1
m
[−Arg(δΣm(r)) + 2πn] .
(34)
Values of n outside the interval [1, m − 1] yield φm,n
coinciding with one of the azimuthal locations inside this
interval.
As pointed out by Bae & Zhu (2018a), these locations
may define additional curves along which constructive
interference occurs.
In their formulation, a spiral arm
may be formed as a result of constructive interference
among the φm,n peaks with different values of m (but
a fixed value of n). The primary spiral is a result of
constructive interference of peaks with n = 0.
In the
inner disk, the secondary spiral is a result of interference
of peaks with n = 1, the tertiary spiral is associated with
n = 2, and so on. In the outer disk, n = m − 1 peaks
are associated with the formation of a secondary spiral,
n = m − 2 with a tertiary spiral, and so on.
The argument of Bae & Zhu (2018a) is based on the
WKB approximation for the phases of the modes, and no
account is given to the behavior of the mode amplitudes.
In our calculations, the mode phases (as well as ampli-
tudes) are computed exactly, by solving for the global
mode structure (i.e., not using WKB approximation).
To illustrate how the mode interference idea works in our
fully self-consistent calculation, in Figure 13 we show the
phases of different crests of the modes defined by equa-
tion (34), which were numerically computed using our
full linear solutions. In agreement with Ogilvie & Lubow
(2002), this figure indicates that the mode phases with
n = 0 are tightly clustered in phase, and these phases
closely follow that of the primary arm. But in addition
to that, one also sees that the phases with n = 1, which
are initially very spread out (over a range of ∼ π), be-
come more clustered towards the inner disk. The phases
of these modes follow the phase of the secondary spi-
ral arm, indicating that the constructive interference of
n = 1 mode crests is indeed responsible for the emergence
of the secondary arm.
The n = 2 phases also start out very spread out in
azimuth at r = rp. They slowly converge as r de-
creases, although they still span a range of ∼ π/2 even at
r = 0.05rp. These phases are approximately coincident
with that of the tertiary arm, although this correspon-
dence is not as tight as it is for the cases of the primary
and secondary arms. This highlights the fact that the
phase information of the modes, while suggestive of the
structure of the spiral arms, is not sufficient to fully cap-
ture their structure. Rather, a full consideration of the
mode phases and amplitudes is required, which is done in
this work.
None of the phases behind the n = 0 phases (i.e. n =
m − 1, n = m − 2) are as tightly clustered as those with
n = 0. Correspondingly, no spiral arm forms behind the
primary in the inner disk (for the fiducial parameters).
15
Fig. 13. -- The phases of different surface density peaks associated with modes with different azimuthal numbers. Each φm,n corresponds
to the nth peak associated with the mode of azimuthal number m. Constructive interference between peaks with different values of m but
the same value of n may be responsible for the formation of the multiple spiral arms. The different colored curves correspond to different
azimuthal numbers, and the different panels correspond to different values of n. In the three leftmost panels, the phases of the n = 0,
n = 1, and n = 2 peaks, which are associated the primary, secondary, and tertiary arms (respectively), are shown, along with symbols
(diamonds) indicating the actual location of these arms. In the two rightmost panels, phases of peaks with n = m − 1 and n = m − 2 are
shown. These peaks are not relevant to secondary spiral arm formation in the inner disk, as they are located behind, rather than in front
of, the primary arm.
5.2. Analytic Estimates
primary arm are then given approximately by
We now try to predict the radius at which secondary
arm emerges, as well as its location and width, using a
phase interference argument. We do this by estimating
the spread in the relevant φm,n's, and identifying the
regions in which they are appropriately clustered as the
locations of the different spiral arms.
In order to do
this, we make use of the WKB approximation, which
we find to reproduce the mode phases (φm,n's) found in
our numerical calculations with reasonable accuracy. For
modes with m close to m∗ ≈ (2hp)−1, the error in the
WKB phases is ∼ 20% in the vicinity of the Lindblad
resonances, and an order of magnitude smaller far from
the resonances, where secondary spiral arm formation
occurs.
From (26) and (34), we have
φm,n = φp− sgn(r− rp)
π
4m
+
2πn
m −
km(r(cid:48))
(cid:48)
dr
(35)
r±
m
(cid:90) r
(cf. Bae & Zhu 2018a). The planet-induced density wake
is dominated by modes with m ≈ m∗. Therefore, defin-
ing φn ≡ φm∗,n, we follow the ansatz of Bae & Zhu
(2018a), and identify in the inner disk the position of
the secondary spiral as φS = φ1, the position of the ter-
tiary spiral as φT = φ2, and so on. The positions of the
secondary and higher-order spiral arms relative to the
φn − φP = −sign(r − rp)
∆km∗ (r(cid:48))
(cid:90) r
−
r±
m∗
+
2πn
m∗
(cid:90) r±
π
4m∗
(cid:48)
dr
+
rp
m∗
km∗ (r(cid:48))
(36)
(cid:48)
,
dr
see equations (32) and (35).
However, in order for a spiral arm to exist at φ =
φn, the appropriate φm,n's must be sufficiently clustered.
The spread in φm,n for values of m in the vicinity of m∗
is
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:18) ∂φm,n
(cid:19)
∂m
m∗
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) ∆m,
δφm,n =
(37)
where ∆m represents the range of azimuthal numbers
which contribute to the arm. We assume that this range
is comparable to the critical m itself, taking ∆m = ζm∗,
with ζ ∼ 1.13 In order for the the φm,n's to construc-
tively interfere and form a spiral arm, we require that
δφm,n < ∆φ0/m∗ (where, as in equation 33, ∆φ0 is a
critical separation required for constructive interference,
13 This is a reasonable assumption for the narrow primary and
secondary spirals. However, the broader widths of the tertiary and
higher order spiral arms may indicate that they are dominated by
modes with a smaller range of m, indicating that ζ < 1 may be
appropriate for these spirals.
−0.50.00.5(φm,0−φp)/π0.10.31.03.0r/rp(a)P−0.50.00.5(φm,1−φp)/π(b)S−0.50.00.5(φm,2−φp)/π(c)T−0.50.00.5(φm,m−1−φp)/π(d)−0.50.00.5(φm,m−2−φp)/π(e)251015202530m16
so that the variation of ψm is less than ∆φ0). Evaluating
the derivative in equation (37) and simplifying, we find
the condition for spiral arm formation
−1m∗δφm,n < ∆ φ0,
δφn ≡ ζ
(38)
where ∆ φ0 = ∆φ0ζ−1 is a new "phase spread" constant
and14
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)2πn − sign(r − rp)
(cid:90) r
r±
π
4
+
Ω2(r(cid:48))dr(cid:48)
s (r(cid:48))km∗ (r(cid:48))
c2
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12) .
δφn =
(39)
Note that δφn represents the spread in mode phases,
while δφn/m∗ gives the approximate spiral arm width
(provided that a spiral arm exists). The integral in equa-
tion (39) is positive (negative) in the outer (inner) disk.
Therefore, a necessary condition for δφn to become small
enough to form a spiral arm is that n ≤ 0 (n ≥ 0) in the
outer (inner) disk. Hence, secondary and higher order
spirals can only be attributed to constructive interference
of φm,n with n > 0 in the inner disk, and only to φm,n
with n < 0 (or equivalently, n = m − 1, m − 2, etc.) in
the outer disk, should spiral arms be present there (Bae
& Zhu 2018a). Note that equation (39) can be used to
estimate the phase spread of the primary spiral arm by
taking n = 0. From this we see that the formation of the
secondary spiral is coincident with the dissolution of the
primary spiral, since δφ0 necessarily becomes large as δφ1
becomes small. This explains a trend seen in Figs. 7g -- i
and 8e -- f: the decrease in δΣP toward the inner disk is
accompanied by an increase in δΣS.
Up to this point, we have left the value of ∆ φ0 arbi-
trary. Roughly speaking, we expect that taking ∆ φ0 ≈ π
in equation (38) should qualitatively predict the presence
of a spiral arm (Ogilvie & Lubow 2002). In practice, we
can use our full numerical results to calibrate this crite-
rion. We compute values of ∆ φ0 corresponding to dif-
ferent relative strengths of the secondary (and tertiary)
spiral arm. We do this by calculating using equation
(39), for example, δφ1(rS20), where rS20 denotes location
at which δΣS/δΣP = 20%, inferred from our numerical
results. These values are given in Table 1.
5.2.1. Passive Spirals
The same formalism can also be applied to our nu-
merical experiments for a passively propagating spirals
presented in §4. In this setup, we assume that all modes
are exactly in phase at rout. We also assume that the
(inner) Lindblad resonances for all modes are located ex-
terior to rout. Therefore, in equation (26), we drop the
π/4 term (associated with the resonance), and take rout
as the lower limit of the km integral. Therefore, we have
φm,n =
2πn
m −
km(r(cid:48))
rout
m
(cid:48)
.
dr
(40)
(cid:90) r
In the expression for the radial wavenumber (28), the or-
bital frequency of the planet Ωp should be replaced by
the specified pattern frequency ωp. Additionally, the ini-
tial width σ of the spiral launched at rout is explicitly
14 In equation (39), n should take on the values 0, ±1, ±2, etc.
While the definition of φm,n (equation 34) also permits values such
as n = m − 1, m − 2, . . ., note that, e.g., φm,m−1 = φm,−1.
specified rather than being set by disk aspect ratio as it
was for a wake launched by a planet. Therefore, we take
m∗ ≈ (2σ)−1 as an approximation of the dominant az-
imuthal mode number in the homogeneous case. We then
have for the arm-to-arm separations and phase spreads
φn − φP =
2πn
m∗ −
∆km∗ (r(cid:48))
(cid:48)
m∗
(cid:90) r
(cid:90) r
For the case ωp = 0, as adopted in §4, the expressions
for φn − φP and δφn take on a simple forms,
and
δφn =
φn − φP =
2πn
m∗
and
where
δφn =
g(x) =
Ω2(r(cid:48))dr(cid:48)
s (r(cid:48))km∗ (r(cid:48))
c2
rout
rout
(cid:12)(cid:12)(cid:12)(cid:12)2πn +
(cid:20)
(cid:12)(cid:12)(cid:12)(cid:12)2πn +
(cid:2)x(q−1)/2 − 1(cid:3)
(cid:26) 2
1 −
m∗
+
q−1
ln(x)
(m2∗ − 1)1/2
dr
,
(cid:12)(cid:12)(cid:12)(cid:12) .
(cid:21) g(r/rout)
(cid:12)(cid:12)(cid:12)(cid:12) ,
hout
(q (cid:54)= 1),
(q = 1).
g(r/rout)
hout(m2∗ − 1)1/2
(41)
(42)
,
(43)
(44)
(45)
5.2.2. Inner/Outer Disk Asymmetry
Our numerical results for planet-driven spirals indicate
that multiple spirals robustly form in the inner disk,
while only a single spiral forms in the outer disk (ex-
cept for small hp, for which a weak secondary spiral is
present). Evidently, this asymmetry is related to the be-
havior of the integral in equation (39) for small versus
large r.
In the outer disk, the integral remains bounded. For
q = 1, its value for r → ∞ is exactly
Ω2(r(cid:48))dr(cid:48)
s (r(cid:48))km(r(cid:48))
c2
= −
2
3hp
ln[m − (m2 − 1)1/2]
(m2 − 1)1/2
.
(46)
(cid:90) ∞
r+
(cid:12)(cid:12)(cid:12)(cid:12) .
(cid:12)(cid:12)(cid:12)(cid:12)2πn −
For q < 1, the value of the integral is even less than given
above, since the integrand is proportional to r(q−6)/2 for
large r. Therefore, for q = 1 and setting m ≈ (2hp)−1
we have
4
3
(47)
ln(hp)
δφn(∞) ≈
π
4 −
Consider δφn with n = −1.
If this quantity becomes
sufficiently small in the outer disk, a secondary spiral
may form. For the fiducial parameters (hp = 0.1),
δφ−1(∞) ≈ 3.95. Since this is larger than the critical
phase spread necessary for secondary spiral arm forma-
tion given in Table 1, no such spiral forms in the outer
disk.
However, equation (47) also indicates that δφ−1 may
become small enough to produce a secondary spiral if hp
is small enough. Indeed, in our numerical calculations,
a weak secondary spiral is found for hp = 0.05. But the
weak (logarithmic) dependence on hp in equation (47)
indicates that δφ−1 can become small enough to produce
a strong secondary spiral only if hp is very small.
can show that(cid:90) r
On the other hand, in the inner disk, for r (cid:28) rp, one
Ω2(r(cid:48))dr(cid:48)
s (r(cid:48))km(r(cid:48)) ≈
c2
g(r/rp) + η
(m2 − 1)1/2hp
,
(48)
r−
where η is an order unity constant. For q ≤ 1, g(r/rp) (cid:29)
1 for small r, and so δφn (equation 39) must cross zero
(indicating constructive interference) before diverging as
r → 0.
In other words, it is always possible to find a
sufficiently small r such that δφn < ∆φ0 (n ≥ 1) for any
∆φ0, and so secondary arm formation is unavoidable in
the inner disk.
5.3. Application of Theory to Our Numerical Results
We now apply analytical results derived above to un-
derstanding certain features of the linear calculations
presented in §3 and §4.
In Figure 14a -- b we display
the theoretical spiral arm phase spreads given by equa-
tion (39) for two different values of hp. The critical radii
rS20, rS50, and rT20 found in our linear calculations are
also indicated. We see that the values of the relevant
δφn's at these critical locations are relatively insensitive
to hp (as well as to the other disk parameters; see Ta-
ble 1). Therefore, equation (38) with properly calibrated
∆ φ0 can be used to reliably predict the location at which
the secondary spiral forms. The phase spreads estimated
from the numerical spiral arm widths (full width at half
maximum for n-th arm, FWHMn), δφn = m∗×FWHMn,
which can be compared to the theoretical δφn's, are also
shown. The theoretical and numerical spiral arm phase
spreads show qualitative agreement. They both vary sim-
ilarly with r, although the exact values differ. This may
be in part due to the the ambiguity in quantifying the
width of a spiral arm found in the numerical calcula-
tion (e.g., using the FWHM instead of some other met-
ric). Figure 14c -- d shows the theoretical arm separations
(equation 36) as well as the arm separations from our
numerical results. The theoretical calculation accurately
predicts the azimuthal separation between the primary
and secondary spirals, but only roughly predicts the sep-
aration of the tertiary and primary spirals. This discrep-
ancy is possibly due to the fact that the tertiary spiral
is not dominated by modes with m ≈ m∗, as assumed in
equation (36), but rather by modes with m (cid:46) m∗.
The theoretical predictions of the critical radii for sec-
ondary spiral formation (for the case of planet-driven
spirals) obtained using equations (38) -- (39) are shown in
Figure 9. These predictions are calibrated using the val-
ues of ∆ φ0 found for the fiducial parameters (i.e.
for
hp = 0.1). That is, the critical values of δφ1 corre-
sponding to different secondary/primary amplitude ra-
tios given in Table 1 for the fiducial parameters are
taken to be universal values applicable for all parame-
ters. These predictions based on simple phase arguments
give excellent agreement with the numerical results over
a range of disk parameters.
The equivalent predictions for the case of passive spi-
rals are shown in Figure 12. These predictions also agree
well with the numerical results. There are, however,
some discrepancies for wide spirals (Fig. 12c). This may
be due to the fact that wide spirals are dominated by a
small number of azimuthal modes, diminishing the ac-
curacy of the approximation (37) for the spread in the
17
TABLE 1
Values of δφn at locations corresponding to several
relative amplitudes of the secondary/tertiary spiral arm
(where for example, rS50 indicates a 50% relative
amplitude of the secondary to the primary).
q
1
1
1
1
0
1
1
hp
0.05
0.07
0.10
0.15
0.10
0.10
0.10
0.10
φm,n's.
1/2
p
1
1
1
1
1
1
0
3/2
δφ1(rS20)
δφ1(rS50)
δφ2(rT20)
2.46
2.39
2.27
2.05
2.19
2.12
2.12
2.36
0.62
0.53
0.62
0.52
0.44
0.28
0.33
0.78
5.75
5.92
5.70
5.08
5.59
5.50
5.43
5.85
Given the success of our analytic arguments in explain-
ing some key features of our numerical calculations, we
can use them to interpret some features of secondary
and higher order spiral arms found in this work. The
weak dependence of the characteristics of multiple arm
on surface density slope p found in both inhomogeneous
(§3) and homogeneous (§4) cases is easily understood in
the context of mode interference. We showed that this
process is well described by the WKB approximation, in
which the mode phases are independent of the Σ profile.
Rather, they depend only on h(r), which in our param-
eterization is fully described by the parameters hp and
q.
The weak remaining dependence on p in the planet-
driven case (see Figure 9c) is likely due to the shift of the
Lindblad resonances at which the modes are launched. In
the homogeneous case, our choice of ωp = 0 effectively
places Lindblad resonances far outside the computation
domain suppressing any sensitivity to p, as stated in §4.
We can also explain why in the homogeneous case, (az-
imuthally) narrower patterns of δΣ produce secondary
arm at smaller radii, see Figure 12c. Narrower perturba-
tions have their power concentrated in modes of higher m
(i.e, they have a larger values of m∗ ≈ (2σ)−1) compared
to azimuthally wider perturbations. Equation (44) then
predicts that δφn becomes small (resulting in phase co-
herence for n = 1, 2 and so on) at lower r/rout as σ is de-
creased (and m∗ is increased correspondingly); note that
g(x) < 0 in the inner disk, see equation (45). Therefore,
narrower perturbations produce high order spiral arms
at smaller radii than the wider ones do.
6. DISCUSSION
The calculations presented in this work (except for
§3.4) are explicitly linear. At the same time, it is well
known that nonlinear effects play an important role in
the propagation and damping of spiral waves, as well as
the evolution of the disk (e.g., Goodman & Rafikov 2001;
Rafikov 2002a, 2016; Arzamasskiy & Rafikov 2018). A
planet-driven spiral wake begins to shock at a distance
of Lsh ∼ (Mp/Mth)−2/5Hp from the planet (in the local
approximation), evolving into a wide "N"-shaped wave
(Goodman & Rafikov 2001). In the process it deposits its
angular momentum into the disk material, so that the an-
gular momentum flux of the wave is no longer conserved
(as it is in linear theory). This process is modified by the
presence of a secondary spiral (Arzamasskiy & Rafikov
2018).
Injection of angular momentum originally car-
ried by the spiral wave into the disk material drives the
18
Fig. 14. -- The secondary/tertiary phase spreads (equation 39; top panels) and positions (equation 36; bottom panels) predicted by
our theoretical phase argument (solid curves) and estimated using numerical calculations (dotted curves). Two different cases are shown:
hp = 0.1 (left) and for hp = 0.05 (right), both with q = p = 1. In the top panels, the dashed vertical lines indicate the radii corresponding
to several critical relative amplitudes of the secondary/tertiary spirals. The values of the theoretical phase spreads at these critical radii,
highlighted by the filled points, are insensitive to the disk parameters. Note that the "reflection" feature of the theoretical δφS in panels
a -- b corresponds to a sign change of the quantity inside the absolute value symbol in equation (39). This feature is not seen in the numerical
δφS, possibly due to the breakdown of the expansion (37) when the spread in the mode phases is small, as a result of the discreteness of m.
evolution of the disk (Arzamasskiy & Rafikov 2018) and
causes gap opening (Rafikov 2002b).
Our linear calculations are strictly valid only as long
as the appearance of the secondary spiral is not preceded
by the shocking of the primary arm. This condition sets
an upper limit on the allowed planet mass. Indeed, if the
secondary spiral emerges after the wake travels a distance
ζHp in the inner disk, then the condition Lsh (cid:38) ζHp
implies that Mp (cid:46) Mthζ−5/2. In our hp = 0.1 calculation
the secondary arm forms at ≈ 5Hp interior to the planet,
meaning that ζ ≈ 5 and Mp (cid:46) 0.02Mth is needed for our
linear calculation to capture the formation of multiple
spirals in quantitative detail. However, at the qualitative
level our calculation should remain valid at substantially
higher values of Mp (e.g., because nonlinear evolution has
only a marginal effect on the analytical phase coherence
calculation presented in §5).
Nonlinear evolution also affects the morphology of the
spirals in the high-Mp regime. Numerical simulations
have shown that the azimuthal separation of the sec-
Mth) planets,
in agreement with our prediction from
ondary and primary arms is ≈ 60◦ for low-mass (Mp (cid:28)
linear theory, but increases up to ≈ 180◦ for massive
(Mp (cid:29) Mth) planets (Dong et al. 2015b; Fung & Dong
2015). This transition is caused by the steady azimuthal
broadening of the spiral wake due to its nonlinear evolu-
tion in the "N-wave" regime (Goodman & Rafikov 2001;
Rafikov 2002a; Zhu et al. 2015). Therefore, the secondary
spirals which we find to form by linear processes should
be regarded just as precursors to the fully-fledged sec-
ondary spiral arms/shocks (see Section 6 of Arzamasskiy
& Rafikov 2018).
In our linear calculations, the ter-
tiary/quaternary arms are always very weak, and so it is
unclear how they are affected by nonlinear effects. How-
ever note that Dong & Fung (2017) reported the pres-
ence of these higher-order arms in nonlinear simulations,
although found them to be destroyed by moderate vis-
cosity.
We note that the global treatment, i.e., accounting for
the cylindrical geometry (as opposed to the local, shear-
ing sheet approximation), is critical for capturing the
formation of multiple spiral arms. We find that the dis-
tance from the planet at which the secondary arm forms
(which we have defined as point at which its amplitude
is 10% of the primary), measured in terms of Hp, is a
decreasing function of hp: the secondary arm forms at
a distance of ≈ 5Hp from the planet for hp = 0.1, and
at ≈ 7.5H from the planet for hp = 0.05.
In the lo-
cal (shearing sheet/box) approximation, corresponding
to the limit hp → 0, in which Hp is the characteristic
length scale, this implies that secondary spirals form at
r − rp/Hp → ∞. In other words, higher-order spirals
would not be captured in the shearing sheet approxi-
mation. Furthermore, secondary spirals form almost ex-
clusively in the inner disk, and not in the outer disk,
an asymmetry that cannot arise in the shearing sheet
framework.
0246810δφS/T[rad]hp=0.1(a)δΣS/δΣP=0.2δΣS/δΣP=0.5δΣT/δΣP=0.20246810hp=0.05(b)δΣS/δΣP=0.2δΣS/δΣP=0.5δΣT/δΣP=0.2TheoryδφSδφTNumericalδφSδφT0.050.10.20.51.0r/rp0.00.51.01.52.02.53.0φS/T−φP[rad](c)0.050.10.20.51.0r/rp0.00.51.01.52.02.53.0(d)TheoryφS−φPφT−φPNumericalφS−φPφT−φP6.1. Comparison with Other Work
The numerical calculations presented in this paper
largely follow those of Ogilvie & Lubow (2002), although
we give a more detailed analysis of the results. However,
Ogilvie & Lubow (2002) did not report the presence of
multiple spirals in their calculations. We find two main
reasons for this. First, they only solved for the pertur-
bation structure down to a radius of 0.3rp (unlike our
calculations, which extend to 0.05rp, allowing the sec-
ondary arm to be fully captured). From our Figure 7,
we see that for our fiducial parameters, the secondary
arm is still quite weak at that radius, with an amplitude
about four times smaller than that of the primary arm.
In their Figure 5, the first hint of a secondary arm be-
comes visible near the inner disk edge, however, it went
unnoticed in their discussion. Second, the calculations
by Ogilvie & Lubow (2002) were restricted to the case of
a constant disk aspect ratio, i.e., q = 1 in our notation.
For flared disks, with q < 1, the secondary arm emerges,
and also overtakes the primary in amplitude closer to the
planet (see Figs. 8 and 9), making its presence more ap-
parent. Finally, we note that Ogilvie & Lubow (2002)
did point out that the phases ψm of modes with differ-
ent m eventually diverge (logarithmically for the case
q = 1) towards the inner disk, so that their construc-
tive interference fails resulting in the partial dissolution
of the primary arm. However, they missed the fact that
the same process also results in convergence of ψm to an
integer multiple of 2π (implying the same value of δΣm)
at different azimuthal locations in the disk, giving rise to
higher order spirals (see §5).
Rafikov (2002a) arrived at the one-armed spiral so-
lution using a method different from Ogilvie & Lubow
(2002). In his case the inability to capture the forma-
tion of higher-order arms is likely caused by a certain
assumption used in the derivation of the linear wake
shape, namely the conservation of the Riemann invari-
ant along the characteristics that cross (rather than fol-
low) the wake. Small changes of this invariant at the
wake crossings, neglected in Goodman & Rafikov (2001)
and Rafikov (2002a), could be responsible for the even-
tual emergence of the secondary spiral. This conjecture
is supported by the fact that secondary spiral emerges
closer to the planet in disks with lower hp: the number
of wake crossings by characteristics (per fixed radial in-
terval) grows as hp goes down, facilitating breakdown of
the one-spiral solution.
Some other ideas for the origin of secondary spirals
have been advanced, in particular, nonlinear effects re-
lated to ultraharmonic resonances with the planet (Fung
& Dong 2015). We do not find these explanations persua-
sive as, first, we reproduce multiple spirals in the frame-
work of a purely linear calculation. Second, our calcula-
tions of passive propagation of a wake with ωp = 0 in §4
do not feature any resonances, and yet, they do result in
secondary spirals.
6.2. Applications
A possible connection between the multiple spirals
driven by a planet and the multiple gaps and rings seen
in some protoplanetary disks was suggested by Dong
et al. (2017, 2018) and Bae et al. (2017).
In the pic-
ture put forth by these authors, a low-mass planet in
19
a low-viscosity disk produces multiple spiral arms, each
of which shock, dissipate, and open a gap at some dis-
tance from the planet. The location of the secondary
gap, attributed to the dissipation of the secondary spi-
ral, was given by Dong et al. (2018) as a function of
planet mass and disk thickness. Our linear prediction
for the location at which the secondary spiral forms is
exterior to their predicted gap location for small planet
masses (Mp (cid:46) 0.2Mth). This is consistent with the sce-
nario in which the secondary spiral first forms in a linear
fashion at some distance from the planet, then propa-
gates inwards, evolving nonlinearly, before shocking and
opening a gap. For larger planet masses, Dong et al.
(2018) predict a secondary gap at a location too close to
the planet for a secondary spiral to have formed in linear
theory. In this case, nonlinear effects clearly play a role
not just in the dissipation of the spiral, but also in its
formation.
The multiple spiral features observed in some pro-
toplanetary disks may be produced by planets (Dong
et al. 2015b). The 180◦ separation of these spirals re-
quires massive planets to produce, and so nonlinear ef-
fects cannot be neglected in these cases. We nonetheless
expect the linear mechanism described in this work to
play an important role in providing the conditions neces-
sary for the formation of these structures (Arzamasskiy
& Rafikov 2018).
However, planets are not the only possibility. As we
showed in §4, any spiral arm, regardless of its origin,
inevitably evolves into multiple spirals as it propagates
through the differentially rotating disk (note that the
assumption of a Keplerian profile for Ω(r) is not es-
sential for the arguments advanced in §5). Therefore,
any mechanism capable of producing at least one spi-
ral arm necessarily produces multiple spiral arms. Pos-
sible mechanisms include gravitational instability (e.g.,
Dong et al. 2015a), accretion from an infalling envelope
(Lesur et al. 2015; Hennebelle et al. 2017), shadows/non-
axisymmetric illumination (Montesinos et al. 2016), and
vortices (Paardekooper et al. 2010). We only require that
density waves with a range of azimuthal mode numbers
are excited and that they are at least somewhat in phase
with one another, so that one or a few well-defined spiral
arms (rather than many flocculent spirals) are produced.
7. SUMMARY
We explored the origin of multiple spiral arms, which
are often observed in protoplanetary disks and also
found in numerical simulations of disks with massive per-
turbers. The two-dimensional structure of surface den-
sity perturbations induced by a planet (as well as that
of a passive spiral) was computed using linear theory of
density wave excitation and propagation (Goldreich &
Tremaine 1979) in the low planet mass (low amplitude)
regime.
We find that, in addition to the strong single spiral
arm excited by the planet in agreement with past studies
(Ogilvie & Lubow 2002; Rafikov 2002a), a secondary spi-
ral arm (and often a tertiary arm, quaternary arm, and so
on) robustly forms in inner disk in the linear regime. The
secondary arm first appears at about r = (0.4 − 0.6)rp,
and, though initially weak, becomes stronger and nar-
rower towards the center of the disk, eventually exceeding
strength of primary arm at ≈ 0.1rp. As the primary arm
20
propagates into the inner disk, the angular momentum
flux it carries gets steadily transferred to these higher-
order spiral arms. In the outer disk, we find that a sec-
ondary spiral arm typically does not form, except for the
coldest disk we considered, with hp = 0.05.
We provide analytical arguments extending the rea-
soning of Ogilvie & Lubow (2002), which show that sec-
ondary spiral arms form as a result of the constructive in-
terference among different azimuthal modes in the inner
disk. Our treatment, which implicitly takes into account
the global variation of both the phases and amplitudes of
different linear modes in a self-consistent manner, thus
corroborates the semi-quantitative, phase coherence pic-
ture previously put forth by Bae & Zhu (2018a) in the
WKB limit. The gravitational potential of the planet
does not play a role in this process. Rather, the planet
only seeds the initial perturbation, which then prop-
agates passively and spawns higher-order spiral arms.
This is confirmed by the persistent emergence of mul-
tiple spirals also in our linear calculations of the passive
inward propagation of an imposed spiral wake, free from
the influence of an external potential (following the setup
of Arzamasskiy & Rafikov 2018).
Our results clearly demonstrate that the formation
of secondary spirals is an intrinsically linear process,
which serves as a precursor for subsequent nonlinear
evolution resulting in a formation of multiple shocks in
the disk. These calculations should help us better un-
derstand planet-driven evolution of protoplanetary disk
(Goodman & Rafikov 2001; Rafikov 2002a), including the
formation of multiple gaps in such disks (Bae et al. 2017;
Dong et al. 2017, 2018). We use them to understand the
details of the global distribution of the torque exerted by
an embedded planet on a disk in Miranda & Rafikov (in
prep.).
Financial support for this study has been provided by
NSF via grant AST-1409524 and NASA via grant 15-
XRP15-2-0139. We thank Stephen Lubow and Gordon
Ogilvie for valuable comments. We are grateful to Wing-
Kit Lee for a careful reading of this paper and for a num-
ber of useful suggestions.
Fig. 15. -- Fractional error in the phase gradient before and after
refinement for the m = 10 mode, with the fiducial parameters
hp = 0.1, q = 1 and p = 1.
satisfies any pair of specified boundary conditions at the
boundaries rin and rout, can be expressed as
δhm = a1δh(1)
H + a2δh(2)
H + δhIH,
(A1)
where a1 and a2 are constants determined by the specific
form of the boundary conditions. We choose outgoing
wave boundary conditions, with
(cid:48)
(cid:48)
m(rin) = Cinδhm(rin), δh
δh
m(rout) = Coutδhm(rout),
(cid:18) DS
(cid:19)
rΣk
(A2)
,
(A3)
APPENDIX
NUMERICAL PROCEDURE
Mode Solutions
where
C = ik +
1
2
d
dr
ln
Here we give a detailed description of the numerical
method used for producing solutions of equation (6). We
compute the values of Laplace coefficients, as well as their
derivatives required for equation (6) (by first differenti-
ating the integrand of equation (13) with respect to α)
using numerical quadrature. We remove the corotation
pole (ω = 0) in equation (6) by replacing ω with ω + iδ,
where δ is a small positive real constant (KP93). We
choose δ = 10−6.
The solution method closely follows the technique de-
scribed in detail by KP93, with several small differ-
ences. We first produce two linearly independent so-
lutions, δh(1)
H of the homogeneous version of
equation (6), by integrating outwards from corotation
starting from arbitrary initial values, as well as an inho-
mogeneous solution, δhIH. The desired solution, which
H and δh(2)
representing an outgoing wave in the WKB limit (e.g.,
Tsang & Lai 2008). Here k = (−DS)1/2/cs is the radial
wavenumber. Note that second term in equation (A3),
describing the slow change in amplitude of the wave, was
not included by KP93 at this step (although they took it
into account in an approximate fashion at a later step).
Now that we have specified the boundary conditions, we
have a system of equations which can be solved for a1
and a2:
(cid:20) d
(cid:20) d
dr
dr
δh(1)
H (rin) − Cinδh(1)
H (rin) − Cinδh(2)
+
δh(2)
=CinδhIH(rin) −
d
dr
δhIH(rin),
H (rin)
H (rin)
a1
a2
(A4)
(cid:21)
(cid:21)
-0.04-0.03-0.02-0.010.000.01ǫ10/k10(a)BeforeRefinement-0.04-0.03-0.02-0.010.000.010.10.31.03.0r/rp(b)AfterRefinement21
Fig. 16. -- Two-dimensional map in polar coordinates (r, φ) of the fractional surface density perturbation δΣ/Σ (scaled by the ratio
pM∗), produced using different values of mmax, the number of azimuthal modes
of the planet mass, Mp, to the thermal mass, Mth = h3
synthesized. The case shown here uses the fiducial parameters.
(cid:20) d
(cid:20) d
dr
dr
(cid:21)
(cid:21)
δh(1)
H (rout) − Coutδh(1)
H (rout) − Coutδh(2)
H (rout)
H (rout)
+
δh(2)
a1
a2
(A5)
=CoutδhIH(rout) −
d
dr
δhIH(rout).
As in KP93, once we have found the values of a1 and
a2 and constructed a solution satisfying the boundary
conditions, we generate a new inhomogeneous solution
δhIH using the "correct" values of δh and δh(cid:48) at corota-
tion (i.e., the ones found from the previously constructed
solution). The system of equations (A4 -- A5) is then re-
solved for new values of a1 and a2. This process ensures
that the final solution is robust with respect to the spe-
cific choice of (arbitrary) homogeneous solutions used in
its construction (equation A1), because it tends to reduce
the absolute values of a1 and a2, so that the contributions
to the correct solution from the homogeneous solutions
are small.
As a final refinement step, also described in KP93, we
minimize the amplitude of the oscillations of the "phase
gradient error" near the boundaries. The phase gradient
error, defined as
(cid:48)
= Arg(δh)
− k,
(A6)
serves as a diagnostic of how close the solution is to an
outgoing WKB wave. The phase gradient error of the ini-
tial solution exhibits oscillations, representing contami-
nation by an in incoming wave, near the boundaries. We
seek to minimize the amplitude of these oscillations by
slightly adjusting the constants Cin and Cout (so that
they are no longer given exactly by equation A3) charac-
terizing the boundary conditions. In principle there are
a variety of ways to achieve this (note that the exact pro-
cedure used by KP93 is not specified). We choose to min-
imize the peak-to-peak amplitudes of the phase gradient
error within one local scale height of each boundary:
Xin = [max() − min()]rin<r<rin+H(rin),
Xout = [max() − min()]rout−H(rout)<r<rout .
(A7)
(A8)
Note that we also tried minimizing the derivative of the
phase gradient error (cid:48) at rin and rout, but found that this
was not as effective at getting rid of the oscillations. We
numerically compute the Jacobian describing the deriva-
tives of X = (Xin, Xout) with respect to C = (Cin, Cout),
and use its inverse to perform one step of the secant
method to find the root of X(C). This greatly reduces
the amplitude of the oscillations of the phase gradient
error, see Figure 15 for an example of this process. In
practice this refinement only requires changing the val-
ues of Cin and Cout by a very small amount ((cid:46) 1%).
Also note that this refinement produces only very imper-
ceptible changes in the form of δh(r) and δh(cid:48)(r), but is
potentially important for ensuring that different modes
have the correct phase when the interference of many
modes is considered.
−0.50.00.5(φ−φp)/π0.10.31.03.0r/rp(a)mmax=10−0.50.00.5(φ−φp)/π(b)mmax=20−0.50.00.5(φ−φp)/π(c)mmax=40−0.50.00.5(φ−φp)/π(d)mmax=80−2.0−1.5−1.0−0.50.00.51.01.52.0(Mp/Mth)−1δΣ/Σ22
As noted in §2, we do not include the indirect poten-
tial term δm,1GMpr/r2
p in our calculations. This term
is proportional to r, and so becomes large for r (cid:29) rp,
in contrast to the direct terms (equation 12), which be-
come small for r (cid:29) rp (note that they both become
small for r (cid:28) rp). Therefore, for m (cid:54)= 1, equation (6)
becomes effectively homogeneous for r (cid:29) rp. Our solu-
tion method exploits this fact by using knowledge of the
asymptotic behavior of the homogeneous equation to set
the outer disk boundary condition (and similarly for the
inner boundary condition). However, for m = 1, when
the indirect term is included, its anomalous behavior at
large r calls into question the validity of setting the outer
boundary condition in this way. We nonetheless carried
out several tests in which the indirect potential was in-
cluded and the outer boundary condition was set under
the homogeneous assumption. We verified that including
the the indirect potential in this way only slightly modi-
fies the profile of the density wake in the outer disk, but
does not otherwise affect our main results.
Mode Synthesis
The two-dimensional surface density perturbation
δΣ(r, φ) is synthesized from the mode solutions δhm us-
ing equations (9) and (14). In order to produce an accu-
rate solution, a sufficient number of modes (up to some
mmax) must be used. The solution must be converged
with respect to mmax, i.e., the perturbation structure
should not change as more modes are added. The value
mmax required for convergence of the two-dimensional
surface density is several times larger than the cutoff pa-
rameter mcut ≈ h−1
p . Figure 16 illustrates this point by
revealing spurious features in the distribution of δΣ(r, φ)
for low mmax = (1−2)h−1
(panels a -- b). Therefore, in all
of our calculations we choose mmax ≈ 8mcut, for which
Figure 16d demonstrates convergence.
p
REFERENCES
Arzamasskiy, L., & Rafikov, R. R. 2018, ApJ, 854, 84
Bae, J., & Zhu, Z. 2018a, ApJ, 859, 118
-- . 2018b, ApJ, 859, 119
Bae, J., Zhu, Z., & Hartmann, L. 2017, ApJ, 850, 201
Baruteau, C., & Masset, F. 2008, ApJ, 672, 1054
Benisty, M., Juhasz, A., Boccaletti, A., et al. 2015, A&A, 578, L6
Ben´ıtez-Llambay, P., & Masset, F. S. 2016, ApJS, 223, 11
de Val-Borro, M., Edgar, R. G., Artymowicz, P., et al. 2006,
MNRAS, 370, 529
Dong, R., & Fung, J. 2017, ApJ, 835, 38
Dong, R., Fung, J., & Chiang, E. 2016a, ApJ, 826, 75
Dong, R., Hall, C., Rice, K., & Chiang, E. 2015a, ApJ, 812, L32
Dong, R., Li, S., Chiang, E., & Li, H. 2017, ApJ, 843, 127
-- . 2018, ApJ, 866, 110
Dong, R., Zhu, Z., Fung, J., et al. 2016b, ApJ, 816, L12
Dong, R., Zhu, Z., Rafikov, R. R., & Stone, J. M. 2015b, ApJ,
809, L5
Fung, J., & Dong, R. 2015, ApJ, 815, L21
Garufi, A., Quanz, S. P., Avenhaus, H., et al. 2013, A&A, 560,
A105
Goldreich, P., Goodman, J., & Narayan, R. 1986, MNRAS, 221,
339
Goldreich, P., & Tremaine, S. 1979, ApJ, 233, 857
-- . 1980, ApJ, 241, 425
Goodman, J., & Rafikov, R. R. 2001, ApJ, 552, 793
Grady, C. A., Muto, T., Hashimoto, J., et al. 2013, ApJ, 762, 48
Hennebelle, P., Lesur, G., & Fromang, S. 2017, A&A, 599, A86
Korycansky, D. G., & Pollack, J. B. 1993, Icarus, 102, 150
Lee, W.-K. 2016, ApJ, 832, 166
Lesur, G., Hennebelle, P., & Fromang, S. 2015, A&A, 582, L9
Lin, M.-K. 2015, MNRAS, 448, 3806
Lin, M.-K., & Papaloizou, J. C. B. 2011, MNRAS, 415, 1445
Maire, A.-L., Stolker, T., Messina, S., et al. 2017, A&A, 601, A134
Montesinos, M., Perez, S., Casassus, S., et al. 2016, ApJ, 823, L8
Muto, T., Grady, C. A., Hashimoto, J., et al. 2012, ApJ, 748, L22
Ogilvie, G. I., & Lubow, S. H. 2002, MNRAS, 330, 950
Ostriker, E. C., Shu, F. H., & Adams, F. C. 1992, ApJ, 399, 192
Paardekooper, S.-J., Lesur, G., & Papaloizou, J. C. B. 2010, ApJ,
725, 146
Petrovich, C., & Rafikov, R. R. 2012, ApJ, 758, 33
Rafikov, R. R. 2002a, ApJ, 569, 997
-- . 2002b, ApJ, 572, 566
-- . 2016, ApJ, 831, 122
Rafikov, R. R., & Petrovich, C. 2012, ApJ, 747, 24
Stolker, T., Dominik, C., Avenhaus, H., et al. 2016, A&A, 595,
A113
Tsang, D. 2014, ApJ, 782, 112
Tsang, D., & Lai, D. 2008, MNRAS, 387, 446
Wagner, K., Apai, D., Kasper, M., & Robberto, M. 2015, ApJ,
813, L2
Wagner, K., Dong, R., Sheehan, P., et al. 2018, ApJ, 854, 130
Ward, W. R. 1997, Icarus, 126, 261
Zhu, Z., Dong, R., Stone, J. M., & Rafikov, R. R. 2015, ApJ, 813,
88
|
1801.05781 | 1 | 1801 | 2018-01-17T18:10:16 | The Propitious Role of Solar Energetic Particles in the Origin of Life | [
"astro-ph.EP",
"astro-ph.HE",
"astro-ph.SR",
"physics.bio-ph",
"physics.space-ph"
] | We carry out 3-D numerical simulations to assess the penetration and bombardment effects of Solar Energetic Particles (SEPs), i.e. high-energy particle bursts during large flares and superflares, on ancient and current Mars. We demonstrate that the deposition of SEPs is non-uniform at the planetary surface, and that the corresponding energy flux is lower than other sources postulated to have influenced the origin of life. Nevertheless, SEPs may have been capable of facilitating the synthesis of a wide range of vital organic molecules (e.g. nucleobases and amino acids). Owing to the relatively high efficiency of these pathways, the overall yields might be comparable to (or even exceed) the values predicted for some conventional sources such as electrical discharges and exogenous delivery by meteorites. We also suggest that SEPs could have played a role in enabling the initiation of lightning. A notable corollary of our work is that SEPs may constitute an important mechanism for prebiotic synthesis on exoplanets around M-dwarfs, thereby mitigating the deficiency of biologically active ultraviolet radiation on these planets. Although there are several uncertainties associated with (exo)planetary environments and prebiotic chemical pathways, our study illustrates that SEPs represent a potentially important factor in understanding the origin of life. | astro-ph.EP | astro-ph | Draft version January 18, 2018
Typeset using LATEX twocolumn style in AASTeX61
THE PROPITIOUS ROLE OF SOLAR ENERGETIC PARTICLES IN THE ORIGIN OF LIFE
Manasvi Lingam,1, 2 Chuanfei Dong,3, 4 Xiaohua Fang,5 Bruce M. Jakosky,5 and Abraham Loeb1, 2
1Institute for Theory and Computation, Harvard University, Cambridge MA 02138, USA
2Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA
3Department of Astrophysical Sciences, Princeton University, Princeton, NJ 08544, USA
4Princeton Center for Heliophysics, Princeton Plasma Physics Laboratory, Princeton University, Princeton, NJ 08544, USA
5Laboratory for Atmospheric and Space Physics, University of Colorado Boulder, Boulder, CO 80303, USA
ABSTRACT
We carry out 3-D numerical simulations to assess the penetration and bombardment effects of Solar Energetic
Particles (SEPs), i.e. high-energy particle bursts during large flares and superflares, on ancient and current Mars. We
demonstrate that the deposition of SEPs is non-uniform at the planetary surface, and that the corresponding energy
flux is lower than other sources postulated to have influenced the origin of life. Nevertheless, SEPs may have been
capable of facilitating the synthesis of a wide range of vital organic molecules (e.g. nucleobases and amino acids).
Owing to the relatively high efficiency of these pathways, the overall yields might be comparable to (or even exceed)
the values predicted for some conventional sources such as electrical discharges and exogenous delivery by meteorites.
We also suggest that SEPs could have played a role in enabling the initiation of lightning. A notable corollary of our
work is that SEPs may constitute an important mechanism for prebiotic synthesis on exoplanets around M-dwarfs,
thereby mitigating the deficiency of biologically active ultraviolet radiation on these planets. Although there are several
uncertainties associated with (exo)planetary environments and prebiotic chemical pathways, our study illustrates that
SEPs represent a potentially important factor in understanding the origin of life.
8
1
0
2
n
a
J
7
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
8
7
5
0
.
1
0
8
1
:
v
i
X
r
a
Corresponding author: Manasvi Lingam & Chuanfei Dong
[email protected] & [email protected]
2
1. INTRODUCTION
Our understanding of solar flares and associated phe-
nomena, such as Solar Energetic Particles (SEPs) and
Coronal Mass Ejections (CMEs), has improved greatly
in recent times, both from a theoretical and observa-
tional standpoint (Webb & Howard 2012; Reames 2013;
Priest 2014; Comisso et al. 2016; Benz 2017). These dis-
coveries have been supplemented by a wealth of data
from the Kepler mission, which surveyed ∼ 105 stars
and yielded detailed statistics concerning the frequency
of large flares with energies & 1033 ergs (superflares)
on M-, K-, and G-type stars (Maehara et al. 2012;
Shibayama et al. 2013; Candelaresi et al. 2014).
Hence, there has been a concomitant increase in stud-
ies analyzing the effects of flares and superflares on plan-
ets situated in the habitable zone (HZ) of their host
stars, i.e. the region in which liquid water can theoret-
ically exist on the surface of the planet (Kasting et al.
1993). Since large flares are usually accompanied by
bursts of high-energy radiation and particles, most stud-
ies have highlighted their deleterious effects from the
perspective of habitability. The dangers posed by SEPs
are especially significant since they may cause significant
ozone depletion and biological damage (Dartnell 2011;
Melott & Thomas 2011; Lingam & Loeb 2017a). How-
ever, recent evidence intriguingly suggests that flares
could have played a beneficial role in the origin of life
(abiogenesis) by delivering the requisite energy for the
prebiotic synthesis of organic compounds (Buccino et al.
2007; Airapetian et al. 2016; Nava-Sedeno et al. 2016;
Ranjan et al. 2017a; Lingam & Loeb 2017b,c).
It is therefore the goal of this paper to examine the
potential role of SEPs, specifically energetic protons, in
facilitating surface-based prebiotic chemistry.
In this
work, we shall not tackle non-surficial theories for the
origin of life, such as hydrothermal vents (Martin et al.
2008; Russell et al. 2014), since the SEP fluxes are not
expected to be significant in these environments. We
will focus on ancient and current Mars, primarily mo-
tivated by the fact that the habitability of both en-
vironments has been extensively investigated (McKay
1997; Schulze-Makuch et al. 2008; Westall et al. 2013;
Cockell 2014). Furthermore, some authors have ar-
gued that the biological potential of Noachian Mars
was similar to, or slightly greater than, that of Hadean
Earth (Jakosky & Shock 1998; Nisbet & Sleep 2001;
Gollihar et al. 2014; Benner & Kim 2015) although its
current value is several orders of magnitude smaller
than present-day Earth (Summers et al. 2002); there
also exists a remote possibility that life on Earth
was seeded by Martian ejecta (Davis & McKay 1996;
Mileikowsky et al. 2000). Subsequently, we will gener-
alize our results to encompass exoplanets, with atmo-
spheres resembling early Mars, orbiting low-mass stars
(K- and M-dwarfs).
2. SIMULATION SETUP AND RESULTS
We adopt the neutral atmosphere from the Mars
Global Ionosphere Thermosphere Model (M-GITM)
(Bougher et al. 2015), which is a 3-D whole atmosphere
(ground-to-exobase) code that captures the Martian
lower atmosphere and its thermosphere-ionosphere. It is
particularly noteworthy that M-GITM, unlike previous
General Circulation Models (GCMs), does not rely on
the hydrostatic assumption and can therefore deal with
large vertical velocities (Ridley et al. 2006; Deng et al.
2008). This feature is particularly relevant since ver-
tical winds must be taken into account when dealing
with extreme space weather events, for e.g. CMEs and
heating due to pickup ions (Fang et al. 2013a).
To simulate the "current" Martian atmosphere, we use
Ps = 6 mbar and 1 EUV, while the "ancient" Mars at-
mosphere ∼ 4 Gya is chosen to be Ps ≈ 1 bar and 10
EUV (Ribas et al. 2005; Boesswetter et al. 2010); here,
Ps denotes the surface pressure whereas 1 EUV refers to
the extreme ultraviolet (EUV) flux under current solar
moderate conditions. It should be noted, however, that
the exact value of Ps for Noachian Mars is unknown, and
it could have been lower than 1 bar (Tian et al. 2009); a
more tenous atmosphere would lead to an enhanced SEP
flux at the surface for a given event. Next, we need to
choose a particular SEP event to undertake our simula-
tions. The input spectrum at the top of the atmosphere
is adopted from the energetic January 2005 SEP event,
and corresponds to an energy flux (per unit energy) of
5 × 1011 pr cm−2 MeV−1 at 0.1 MeV, while the associ-
ated spectral index is −2.15 (Mewaldt et al. 2012). We
note that the 2005 SEP event has been studied previ-
ously as a representative example of a high-fluence solar
proton event (associated with an X7 solar flare) in the
context of prebiotic chemistry (Airapetian et al. 2016).
SEP events with total fluences greater than the January
2005 event have been documented (Mewaldt 2006), and
hence it might constitute a representative example for
the active young Sun. However, it must be recognized
that the SEP spectrum and initial flare energy are not
necessarily correlated, since the causal mechanisms for
SEPs are complex and varied (Reames 2013).
To calculate the penetration of SEPs through the at-
mosphere, we apply a continuously slowing down ap-
proximation (CSDA) energy loss model that was devel-
oped by Jackman et al. (1980) and recently improved
by Fang et al. (2013b). The model has been extensively
validated for energetic particle transport, showing ex-
cellent agreement with results from more sophisticated
collision-by-collision calculations of the combined pri-
mary ion effects (Fang et al. 2004; Jolitz et al. 2017) and
secondary electron effects (Lummerzheim et al. 1989).
Due to the lack of reliable collisional cross-section mea-
surements, the collision-by-collision method cannot be
extended to high-energy particles beyond ∼ 1 MeV. In
contrast, the stopping power data published by the Na-
tional Institute of Standards and Technology covers a
broader energy range, up to 10 GeV (Berger et al. 2011).
3
Figure 1. Panel (a) illustrates the globally averaged differential number flux distribution at the planetary surface. Different
colors represent cases with varying cutoffs marked at the top. The black dashed line shows the precipitating SEP energy spectral
shape for reference. Panel (b) depicts the altitude profiles of penetrating SEP energy fluxes near the subsolar point under current
(in black) and ancient (in red) atmospheric conditions. The energy fluxes are divided according to different energy ranges of
incident particles at the topside boundary, as marked in the figure. Panel (c) presents the globally averaged ionization altitude
profiles for < 1 GeV and < 6 GeV SEP precipitation in the current (black) and ancient (red) Martian atmosphere respectively.
The shaded regions demarcate the range of values at different latitudes and longitudes.
In Fig. 1(a), the penetrating SEP energy spectrum at
the surface is shown for current and ancient Mars. From
Fig. 1(b), it is seen that SEPs with initial energy lower
than ∼ 150 MeV (black dotted and dashed lines) are
unable to penetrate through the current Martian atmo-
sphere, and only those energetic particles with energies
& 150 MeV (black dashed-dotted line) can reach the
surface; the same cutoff value has also been presented
4
Figure 2. Atmospheric column mass density and penetrating SEP flux at the surface for current (< 1 GeV) and ancient (< 6
GeV) Mars as a function of latitude and local time.
by the Mars Science Laboratory's Curiosity rover group
(Hassler et al. 2014). In contrast, the much thicker at-
mosphere assumed in the ancient epoch effectively pre-
vents the penetration of . 150 MeV particles above 40
km altitude (red dotted line). In fact, the threshold en-
ergy for incident SEPs being able to reach the surface
through the ancient Martian atmosphere is elevated to
∼ 3 GeV. Note that the threshold penetration energies
of ∼ 150 MeV and ∼ 3 GeV, for current and ancient
Mars respectively, vary slightly over the planet due to
the spatial asymmetry of the atmospheres. The inci-
dent SEP spectrum at the top of the atmosphere is also
shown for reference in Fig. 1(a).
In Fig. 2, column mass density and SEP energy de-
position on the planetary surface as a function of lat-
itude and longitude is presented. This is one of the
notable features of our study since our simulations are
fully 3-D and thus display signatures of asymmetry
(Bougher et al. 2006, 2015). For current and ancient
Mars, the SEP energy flux broadly increases as one
moves to higher latitudes. Moreover, for both cases, the
energy flux is generally higher on the night-side relative
to the day-side. The temperature difference between the
day- and night-side affects the global neutral density dis-
tribution - see column mass density in Fig. 2 - and con-
sequently the SEP energy deposition at the planetary
surface is altered, thus leading to the observed asymme-
try. These asymmetric features are less pronounced for
ancient Mars due to significant day-to-night transport
caused by the strong day-side EUV heating; the trans-
port leads to a relatively uniform distribution of the at-
mosphere around Mars. The resulting consequences are
explored further in Sec. 3.3.
We have neglected the effects of weak planetary mag-
netic fields on SEPs in our simulations. Once the Mar-
tian dynamo stopped functioning ∼ 4.1 Gya, the mag-
netic field strength is believed to have declined rapidly
(Lillis et al. 2008; Fassett & Head 2011; Lillis et al.
2013); note, however, that a later age has been pro-
posed for the shutdown of the Martian dynamo by
some authors (Schubert et al. 2000). The high energy
of SEPs in conjunction with weak magnetic fields result
in a very large gyroradius that may even exceed the
planetary scale, suggesting that the deflection of SEPs
by weak magnetic fields is minimal. Exoplanets around
M-dwarfs could also be characterized by weak magnetic
fields (Shields et al. 2016), and will be discussed later
in Sec. 3.4.
3. IMPLICATIONS FOR PREBIOTIC CHEMISTRY
We study some of the ensuing implications of the
above results for Noachian Mars and other exoplanets
(with similar atmospheres) in the HZ of their host stars.
3.1. Energy source for prebiotic chemistry
The origin of life required suitable energy sources for
the synthesis of prebiotic compounds (Ehrenfreund et al.
2002; Morowitz & Smith 2007; Luisi 2016; Walker
2017). Commonly studied energy pathways in this
regard include solar radiation,
shock heating from
impacts, electrical discharges,
radioactivity, volcan-
ism and geochemical energy (Miller & Urey 1959;
Maher & Stevenson 1988; Chyba & Sagan 1992; Pascal
2012; Ruiz-Mirazo et al. 2014).
It is therefore advan-
tageous to estimate the energy available through SEPs
that reach the surface and compare them against the
aforementioned sources.
The energy flux ΦE (in units of J m−2 s−1) can be
estimated as follows:
ΦE = N φE,
(1)
where φE is the globally averaged energy per unit area
deposited on the planetary surface (during a characteris-
tic SEP event) and N is the number of such SEP events
per day.1 Our simulations yield φE = 1.5 × 10−1 J/m2
for ancient Mars, and we must now determine the value
of N . The study of ∼ 105 G-type stars by Kepler yielded
a power-law distribution for the occurrence frequency of
superflares with a spectral index α between −1.5 and
1 In reality, ΦE is time-averaged, and it is higher during the
transient flaring period.
5
−2.3 (Maehara et al. 2012, 2015). It has been suggested
that superflares can occur on the Sun (Shibata et al.
2013; Mekhaldi et al. 2015), although the evidence re-
mains disputed (Schrijver et al. 2012). A similar power-
law scaling has been inferred for regular flares on the
Sun, but the spectral index is slightly different when
compared to the corresponding estimate for superflares
(Crosby et al. 1993; Hannah et al. 2011).
As we are concerned with the young Sun, expected to
have been more active ∼ 4.0 Gya (Gudel 2007), it seems
reasonable to assume that it obeyed statistics similar to
active G-type stars; the latter are predicted to have a su-
perflare occurrence rate of 0.1 per day (Shibayama et al.
2013). However, we caution that the young Sun was not
necessarily characterized by statistics similar to active
solar-type stars studied by the Kepler mission. With
this caveat in mind, we use the above value to con-
clude that the frequency N0 of Carrington-type events
for the ancient young Sun may have been N0 . 40
per day, which is lower by a factor of ∼ 6 compared
to Airapetian et al. (2016); these events are less pow-
erful than superflares but occur more frequently with
adequate energy release. Although large flares are often
accompanied by CMEs and SEPs (Emslie et al. 2012;
Reames 2013; Desai & Giacalone 2016), not every such
event will impact the planet. The number of impacting
events is approximately estimated to be
N = N0 sin2(cid:18) θ
2(cid:19) ,
(2)
0 sin θ′dθ′R 2π
cles, i.e. Ω/4π where we have Ω = R θ
where the second factor is purely geometric, and is based
on non-isotropic emission with an opening angle θ. It
stems from the solid angle fraction of the emitted parti-
0 dφ′ =
2π (1 − cos θ). The opening angle ranges between 20◦
and 120◦ and we select a fiducial value of θ ∼ 47◦ based
on observational evidence (Yashiro et al. 2004). Upon
substituting this value into (2), we find N . 6 events per
day. In reality, N will be even lower since other factors
(e.g. magnetic field orientation) should be taken into
consideration (Gopalswamy et al. 2007). We will there-
fore normalize N by 1 event per day; see also Odert et al.
(2017) in this context. Using this data in (1), we find
ΦE = 0.05 f (cid:18) N
1 day−1(cid:19) kJ m−2 yr−1,
(3)
where we have introduced the enhancement factor f due
to the following reason. We have implicitly assumed that
only Carrington-type flares contribute to solar proton
events impacting the planet. However, in reality, super-
Carrington events with flare energies & 1033 ergs are
expected to exist, as mentioned earlier (Shibata et al.
2013). Although these events are rarer, they would be
characterized by higher SEP fluences (Takahashi et al.
2016). Thus, f & 1 serves as the enhancement fac-
tor arising from these events, and is determined from
6
the ratio of the cumulative and probability distribution
functions describing the frequency of flares (with a given
energy), and will depend on both α and the energy cutoff
(Newman 2005). We will normalize f by unity hence-
forth, since f = 1 constitutes the lower bound.
The value of ΦE in (3) is lower than some energy
fluxes on ancient Earth by 1-2 orders of magnitude
(Deamer & Weber 2010), such as electrical discharges
(2.9 kJ m−2 yr−1)
and volcanism (5.4 kJ m−2 yr−1).
However, it is worth noting that other energy sources
are much higher, for e.g. the energy flux from solar UV
radiation is about 7 orders of magnitude greater.
It
must also be pointed out that the energy flux due to
volcanism might have been much higher on early Mars
(Halevy & Head 2014). Here, we have compared ΦE for
Noachian Mars with other energy sources on Hadean-
Archean Earth, since a comprehensive knowledge of
the corresponding energy fluxes for ancient Mars is cur-
rently lacking. However, in light of the many unknowns,
the above estimates should not be perceived as being
definitive.
E
& ΦMars
E
We can conjecture that ΦEarth
because: (i)
the SEP fluence falls off with the square of the distance
(Feynman et al. 1993; Cooper et al. 2003), and (ii) the
surface pressures and atmospheric compositions of an-
cient Mars and Earth appear to have been fairly similar
(McKay 2010; Arndt & Nisbet 2012; Wordsworth 2016).
Since the SEP energy flux at the planetary surface de-
pends on the incident SEP fluence and atmospheric
properties, (i) and (ii) would therefore collectively imply
that the SEP energy flux for Hadean-Archean Earth was
higher than, or comparable to, that of Noachian Mars.
Hence, the same conclusions discussed previously in the
context of ΦMars
, such
as the comparison of the SEP energy flux against other
sources (e.g. volcanism and lightning).
would also be valid for ΦEarth
E
E
It is possible to compute the number flux ΦN of SEPs
through a similar procedure. Thus, we estimate ΦN via
ΦN = f N φN ,
(4)
where f is the enhancement factor introduced earlier, N
is the number of Carrington-type SEP events per day,
and φN ∼ 7.5 × 108 cm−2 from our simulations. Upon
substituting these values into (4), we end up with
ΦN = 8 × 103 f (cid:18) N
1 day−1(cid:19) cm−2 s−1,
(5)
and further implications are discussed in Sec. 3.3.
3.2. Chemical pathways for prebiotic synthesis
Previously, we have argued that SEPs provide a valu-
able source of energy for prebiotic synthesis and that
their energy flux is lower compared to lightning and vol-
canism. It is, however, important not only to evaluate
the energy fluxes from different sources but also their
efficiency in terms of chemical synthesis (Deamer 1997).
SEPs are known to facilitate the formation of nitro-
gen oxides, for e.g. nitric oxide (NO) and nitrogen
dioxide (NO2), by reacting with atmospheric nitro-
gen (Crutzen et al. 1975; L´opez-Puertas et al. 2005).2
It was noted in Airapetian et al. (2016) that these
molecules reacted with the chemical species CH as fol-
lows:
NO + CH → HCN + O,
N2O + CH → HCN + NO,
(6)
thereby resulting in the production of hydrogen cyanide
(HCN). In addition, other pathways besides (6) lead-
ing to SEP-driven HCN formation were also iden-
tified.
HCN is important in prebiotic chemistry
(Ferris & Hagan 1984) because it is required for the
only known pathways leading to the prebiotic synthe-
sis of (i) nucleic acids, (ii) proteins, and (iii) lipids
through reductive homologation (Patel et al. 2015); in
turn, these molecules are postulated to have been impor-
tant components of protocells. The chemical reactions
leading to (i)-(iii) occur in the presence of UV radiation
with hydrogen sulphide (H2S) serving as the reductant.
It has therefore been argued that HCN is an impor-
tant "feedstock" molecule that may play a vital role in
abiogenesis (Sutherland 2016). Hence, SEPs (and other
energy sources) enable the production of HCN, which
can undergo further UV-mediated chemical reactions to
generate the above prebiotic compounds. However, a
possible limitation is that HCN needs to be transported
from the atmosphere to the surface, where the pathways
of Patel et al. (2015) are functional.
Most of the basic ingredients presumably required for
SEP-driven prebiotic synthesis, such as N2, CO, CO2,
H2O, H2S and CH4, are expected to have been present
on Noachian Mars, although the exact composition re-
mains very uncertain (Owen 1992; Jakosky & Phillips
2001; Farquhar et al. 2000; Formisano et al. 2004;
Halevy et al. 2007; Webster et al. 2015; Wordsworth
2016); these ingredients were also potentially available
on ancient Earth for prebiotic synthesis (Zahnle et al.
2010; Arndt & Nisbet 2012). Moreover, the bioactive
UV flux ∼ 4 Gya emitted by the Sun is predicted to have
been a few times times higher than its present-day value
(Rugheimer et al. 2015; Rapf & Vaida 2016), which can
enable independent UV-mediated prebiotic chemistry
(Ranjan & Sasselov 2017; Ranjan et al. 2017b). Thus,
it appears plausible that prebiotic pathways driven by
SEPs (and UV radiation) were functional on ancient
Mars and Earth, consequently rendering both planets
conducive to the origin of life in this particular scenario.
2 Nitrous oxide (if produced) is a potent greenhouse gas
(Canfield et al. 2010; Airapetian et al. 2016), and may therefore
partly provide a resolution of the "faint young Sun" paradox
(Sagan & Mullen 1972; Kasting 2010; Feulner 2012).
In addition, gaseous mixtures - comprising of CO,
CO2, N2 and H2O - have also been shown in the labo-
ratory (i.e., under controlled conditions) to yield a wide
range of organic compounds when subjected to irradia-
tion by energetic protons with energies greater than a
few MeV. The organic molecules thus produced in the
laboratory could, in principle, also be directly synthe-
sized by SEPs. A few examples of the salient organic
compounds synthesized through this process include the
following:
• Uracil, guanine, adenine and cytosine (Kobayashi et al.
1990; Miyakawa et al. 2002b), which represent
four of the five nucleobases of RNA and DNA.
The absence of thymine (DNA only) might fa-
vor the emergence of RNA prior to DNA, in ac-
cordance with the RNA world hypothesis (Joyce
2002; Orgel 2004). In contrast, meteorites ostensi-
bly lack cytosine as well as thymine, thus making
the eventual synthesis of both RNA and DNA
(via this route) rather difficult (Pearce & Pudritz
2016).
• Amino acids such as alanine, aspartic acid, glycine
and serine (Kobayashi et al. 1998; Miyakawa et al.
2002b); most of these biomolecules are considered
essential
for protein synthesis (Weber & Miller
1981; Zaia et al. 2008; Higgs & Pudritz 2009) and
some authors have suggested that the co-evolution
of amino acids and nucleic acids might have oc-
curred (Koonin & Novozhilov 2009).
Some of
these proteinogenic amino acids have also been
produced by means of UV-mediated chemical
pathways starting from HCN, as noted previously
(McCollom 2013; Sutherland 2017).
• Aromatic compounds, e.g. imidazole (Kobayashi et al.
1995), which could play an important role in abio-
genesis (Ehrenfreund et al. 2006).
• Hydrogen cyanide,
formaldehyde and complex
amino acid precursors (Kobayashi et al. 1998).
Furthermore, if ammonia is unavailable in primordial
planetary atmospheres, it has been suggested that ir-
radiation by high-energy protons would be an impor-
tant pathway for the synthesis of amino acid precursors
(Kobayashi et al. 2001). Lastly, the G-values associated
with amino acid synthesis through irradiation by ener-
getic protons (∼ 0.01) are amongst the highest docu-
mented for prebiotic pathways (Kobayashi et al. 1990).
This is an important point since it indicates that, de-
spite the lower energy flux of SEPs, the resultant yield
of organic molecules can be higher when compared to
a few other energy sources. We will further quantify
this statement by focusing on proteinogenic amino acids
and nucleobases, i.e. the building blocks of proteins and
nucleic acids respectively.
7
However, before proceeding further, a few caveats are
worth noting here. Given the significant uncertainties
concerning the state of early Earth's atmosphere(s), the
gaseous mixtures considered in the above papers could
have been more reducing than that of the Hadean-
Archean Earth. Second, the yields are dependent on
the atmospheric composition, i.e. on the partial pres-
sure of CO. Lastly, the experiments carried out relied
on irradiation by protons of ∼ 3 MeV, whereas the en-
ergies of SEPs can attain maximum values of a few GeV.
As there is no experimental evidence available currently
concerning prebiotic synthesis by particles at these en-
ergies, we will henceforth operate under the assumption
that protons of different energies impact prebiotic chem-
istry uniformly; a similar assumption has also been in-
voked in analyzing endergonic chemical reactions driven
by cosmic rays (Miyakawa et al. 2002b). Although such
an assumption will be utilized for carrying out quanti-
tative estimates, we emphasize that there is a pressing
need for follow-up experiments to address this important
question.
The rate of amino acids synthesized, denoted by
MA,
is proportional to the deposited energy flux
(Kobayashi et al. 1995). Using the G-values provided
in Kobayashi et al. (1990) and Miyakawa et al. (2002b),
we obtain the following phenomenological relation
MA = 107 kg/yr (cid:18)
ΦE
0.1 kJ m−2 yr−1(cid:19) ,
(7)
for Earth-sized planets, where we have substituted the
characteristic amino acid molar mass of 0.1 kg/mol. The
above formula follows from multiplying the energy flux
with the efficiency of amino acid synthesis. This value is
comparable to the yields of organic molecules from other
energy sources as seen from Fig. 3 of Deamer (1997).
However, it should be noted that (7) deals solely with
the synthesis of amino acids, whereas the latter (Deamer
1997) depicted the estimates for all organic compounds.
In a similar fashion, the amount of nucleobases produced
can be evaluated accordingly. We end up with
ΦE
0.1 kJ m−2 yr−1(cid:19) ,
(8)
MN = 104 kg/yr (cid:18)
MN denotes the rate of nucleobases synthesized,
where
and we have used the fact that the typical molar mass
of nucleobases is 0.1 kg/mol.
As a point of reference,
let us consider the ex-
ogenous delivery of nucleobases and amino acids by
meteorites.
intact or-
ganics from Chyba & Sagan (1992) along with data
from carbonaceous chondrites (Kvenvolden et al. 1970;
Cronin & Pizzarello 1983; Pizzarello & Shock 2017),
MA ∼ 300 kg/yr and MN ∼ 2 kg/yr
we arrive at
(Miyakawa et al. 2002b). Thus, provided that the con-
stituent gases are available in sufficient amounts, prebi-
otic synthesis of amino acids and nucleobases by SEPs
Using the delivery rate of
8
could have been approximately four orders of magni-
tude higher than delivery by meteorites; this follows
from comparing the above values for meteorites with
(7) and (8).
For weakly reducing atmospheres we find that the
yield of amino acids via electrical discharges is ∼ 5 × 107
kg/yr using the data from Stribling & Miller (1987)
and Miller (1998). A photochemical model was em-
ployed by Tian et al. (2011) to compute the surface de-
position rates of HCN for different concentrations of
methane, evaluated at different levels of carbon diox-
ide. Prior to the advent of methanogens, it was con-
cluded that the prebiotic deposition rate of HCN was
∼ 107 molecules cm−2 s−1. Upon converting this into
kg/yr and using the conversion efficiency of HCN to
amino acids (Stribling & Miller 1987), we find that the
yield of amino acids would be ∼ 3 × 107 kg/yr. The
values for electrical discharges and UV photochemistry
are commensurate with the amount of amino acids that
could be produced by means of SEP-mediated synthe-
sis, as seen from (7). In light of the many uncertainties
surrounding prebiotic pathways and the composition of
early Earth/Mars atmospheres, we emphasize that all
of the preceding estimates should be regarded as being
heuristic.
These organic compounds, especially nucleobases but
also amino acids, will be deposited at very low con-
centrations on the planetary surface (land and oceans)
and subsequent prebiotic chemistry would face further
challenges (Budin & Szostak 2010).3
In order for nu-
cleotides to undergo rapid polymerization and yield
nucleic acids, wet-dry cycles and thermal gradients are
expected to play an important role on thermodynamic
grounds via polymerase chain reactions (Kreysing et al.
2015; Ross & Deamer 2016).4 A wide range of pu-
tative environments endowed with operational cycles
(and gradients) have been identified on Earth,
for
e.g.
intermountain valleys and
beaches (Bywater & Conde-Frieboesk 2005; Adam 2007;
Benner et al. 2012; Da Silva et al. 2015; Lingam & Loeb
2017d) to name a few. Most of these habitats also
supply the requisite minerals (and nutrients) that can
play an important role in abiogenesis by catalyzing
hydrothermal pools,
3 The putative role of aerosols in facilitating abiogenesis merits
consideration (Tuck 2002; Donaldson et al. 2004) in connection
with the above limitation, since they function as non-equilibrium
chemical reactors and can drive the concentration of reactants
(Dobson et al. 2000; Stueken et al. 2013).
4 Alternatively, on a cold and wet ancient Mars (Fair´en 2010;
Grott et al. 2011) or Earth (Zahnle et al. 2010), abiogenesis might
have been facilitated through the concentration of prebiotic com-
pounds by means of eutectic freezing (Miyakawa et al. 2002a;
Bada 2004; Price 2007; Monnard & Szostak 2008). Ice and freeze-
thaw cycles may have also played an important role in driving
the assembly of RNA polymerase ribozymes (Trinks et al. 2005;
Bartels-Rausch et al. 2012; Attwater et al. 2013; Mutschler et al.
2015).
polymerization (Ferris et al. 1996; Hazen & Sverjensky
2010; Cleaves II et al. 2012) and enabling homochirality
(Hazen & Sholl 2003; Lambert 2008).
However, it remains quite ambiguous as to whether
analogous environments and mechanisms could have
been prevalent on Noachian Mars (Carr & Head 2010).
Surface water, minerals, and (perhaps) oceans are
known to have existed intermittently on ancient
Mars (Baker 2001; Solomon et al. 2005; Tosca & Knoll
2009; Di Achille & Hynek 2010; Adcock et al. 2013;
Ehlmann & Edwards 2014). Conversely, geological ex-
plorations of Meridiani Planum by the Mars Exploration
Rover Opportunity suggest that the Martian paleoen-
vironment was likely to have been acidic, arid and ox-
idizing (Squyres & Knoll 2005; Hurowitz & McLennan
2007; Ehlmann et al. 2011), while the salinity,
ionic
strength and chaotropic activity were potentially higher
than the tolerance levels of current terrestrial or-
ganisms (Tosca et al. 2008; Ball & Hallsworth 2015;
Fox-Powell et al. 2016). These factors may have collec-
tively posed difficulties for subsequent prebiotic chemi-
cal reactions and abiogenesis to occur on ancient Mars
(Knoll et al. 2005), although the environmental con-
ditions at the Endeavour and Gale craters during the
Noachian and/or Hesperian eras were possibly more fa-
vorable for the emergence of life (Arvidson et al. 2014;
Hurowitz et al. 2017).
3.3. Other implications of SEPs
Previously, we have outlined how SEPs may provide
a direct energy source for prebiotic synthesis. However,
there are other avenues by which they can indirectly
contribute to the latter as well.
Physical mechanisms have been proposed wherein rel-
ativistic runaway electron avalanches (RREAs) associ-
ated with the formation of cosmic ray secondaries may
lead to the initiation of lightning (Gurevich et al. 1992).
The basic principle behind RREAs is that runaway
electrons can undergo Møller scattering and give rise
to other free electrons which also exceed the runaway
threshold, thereby giving rise to a cascade. However,
in order to trigger RREAs, "seed" particles with suf-
ficient energy are required.
It has been hypothesized
that these seed particles were primarily contributed by
cosmic ray secondaries (Gurevich et al. 1999). As SEPs
can reach energies of a few GeV, it seems plausible that
they could also serve as the seed particles for RREAs.
There exists some recent observational evidence in this
regard favoring positive correlations between elevated
levels of SEPs (and solar activity) and lightning rates
(Siingh et al. 2011; Scott et al. 2014). If this conjecture
were valid, higher stellar activity would lead to enhanced
lightning activity on (exo)planets (Hodos´an et al. 2016).
The relativistic runaway breakdown theory predicts
that the flux of runaway electrons ΦRE is linearly pro-
portional to the flux of energetic seed particles ΦN
(Dwyer & Uman 2014), and the latter is given by (5).
The SEP number flux reaching the surface was prob-
ably comparable to the Galactic Cosmic Ray (GCR)
flux (Sanuki et al. 2000).
If this were correct, solar
proton events on ancient Mars (and Earth) may have
contributed significantly to ΦRE implying that SEPs,
and therefore flares, could have played an important
role in the initiation of lightning on these planets. The
relevance of electrical discharges for prebiotic chem-
istry has been comprehensively studied (Miller 1953;
Miller & Urey 1959; Borucki & Chameides 1984; Bada
2013; McCollom 2013), especially in the context of ni-
trogen fixation; the latter's corresponding budget for an-
cient Mars is predicted to have been similar to that of
early Earth (Segura & Navarro-Gonz´alez 2005).
From Fig. 2, the higher latitudes and night-side are
characterized by slightly increased SEP energy fluxes at
the surface for both current and ancient Mars. Since
MA,N ∝ φE follows upon combining (1), (7) and (8),
the synthesis of organic compounds could be higher at
the above regions (up to a factor of ∼ 2 for current
Mars). Thus, the possibility that favorable conditions
for abiogenesis existed at higher latitudes on ancient
Mars, Earth and other exoplanets may merit consider-
ation (Jakosky et al. 2003), although it might be coun-
terbalanced by the decline in reaction rates at relatively
lower temperatures (Lingam & Loeb 2017e).
Hitherto, we have discussed the positive implica-
tions of SEPs, but we wish to reiterate that they
can also be detrimental to life in surface environ-
ments. The production of hydrogen and nitrogen oxides
by SEPs rapidly depletes the ozone layer (Solomon
1999; L´opez-Puertas et al. 2005),
thereby enabling
harmful UV-B and UV-C radiation to reach the sur-
face and potentially triggering biological extinctions
(Melott & Thomas 2011; Lingam & Loeb 2017a). SEPs
could also trigger the formation of secondary energetic
particles, and lead to high radiation doses, especially on
planets in the HZ of low-mass stars, which can prove to
be detrimental to certain complex lifeforms on the sur-
face (Atri 2017; Tilley et al. 2017). On the other hand,
surficial organisms may adapt to such high-radiation en-
vironments by means of ultraviolet screening compounds
(Cockell & Knowland 1999). Furthermore, habitats in
the deep biosphere or the oceans (Cleaves & Miller
1998) should be well-suited for protecting biota from
ionizing radiation.
In addition, extreme space weather events also en-
hance atmospheric loss due to erosion by the so-
lar/stellar wind (Dong et al. 2014, 2015a,b, 2017a, 2018;
Jakosky et al. 2015; Ma et al. 2017).
if the
atmosphere is altogether depleted in . O(100) Myr
(Dong et al. 2017b; Lingam & Loeb 2017e), there might
not be sufficient time for life to originate and evolve
(Lingam & Loeb 2017f,b); see Spiegel & Turner (2012)
for a detailed Bayesian discussion of the constraints on
the characteristic timescale for abiogenesis.
In fact,
9
3.4. Implications of SEPs for exoplanets
Let us now turn our attention to exoplanets in the HZ
of other stars. We will focus primarily on M-dwarfs, as
they are numerous in our Galaxy and detecting exoplan-
ets around them is comparatively easier.
Using the inverse-square scaling of SEP fluence
(Feynman et al. 1993) with orbital distance a, we have
φN ,E ∝ a−2 and the available empirical evidence indi-
cates that this approximation may be reasonably valid
for exoplanets in the HZ of M-dwarfs (Youngblood et al.
2017). Assuming that the atmospheric properties are
akin to that of ancient Mars, we obtain
ΦE = 50 f (cid:18) N
1 day−1(cid:19)(cid:16)
−2
a
0.05 AU(cid:17)
kJ m−2 yr−1, (9)
and the last factor on the RHS accounts for the closer
star-planet distance. We have retained the same nor-
malization factor for N given that M-dwarf exoplanets
are impacted by . 5 events per day (Kay et al. 2016).
Similarly, we find that the number flux scales as
ΦN = 7.3 × 106 f (cid:18) N
1 day−1(cid:19)(cid:16)
a
0.05 AU(cid:17)
−2
cm−2 s−1.
(10)
From (9) and (10), we see that the energy and number
fluxes for exoplanets in the HZ of low-mass M-dwarfs
are likely to be ∼ 100 − 1000 times higher than that of
ancient Mars (or Earth). In turn, this has a number of
consequences delineated below.
• The energy flux is expected to be higher than the
corresponding values for most other sources on
early Earth, with the exception of UV radiation.
It can, for instance, exceed the energy fluxes for
shock impacts and radioactivity.
• The number flux ΦN is likely to be higher than
that of GCRs by several orders of magnitude. In
turn, SEPs could prove to be a fairly major player
in the initiation of lightning.
• From (7) and (8), we see that
MA,N ∝ ΦE.
Hence, SEPs might serve as the most significant
mechanism for the synthesis of amino acids and
RNA/DNA nucleobases on exoplanets around M-
dwarfs; we find that ∼ 106 − 107 kg/yr of nucle-
obases might be produced via SEPs and this value
is comparable to, or slightly higher than, the up-
per bound of ∼ 106 kg/yr estimated from Table 1
of Nava-Sedeno et al. (2016).
• The above point is rendered even more impor-
tant by the fact that UV-mediated prebiotic path-
ways are potentially inefficient on M-dwarf ex-
oplanets, owing to the lower bioactive UV flu-
ence (by a factor of ∼ 1000) reaching the surface
(Buccino et al. 2007; Rugheimer et al. 2015). It is,
10
however, quite possible that flares may ameliorate
this UV deficiency to some degree (Smith et al.
2004; Buccino et al. 2007; Ranjan et al. 2017a).
• Since ΦE is proportional to N , as seen from (9),
it could imply that stars with higher flare activ-
ity are more conducive to prebiotic synthesis via
SEPs (and even UV radiation). However, the same
factors can also prove to be detrimental to the sus-
tenance of complex biospheres (Dartnell 2011).
Most of the above conclusions concerning M-dwarf ex-
oplanets are also likely to be applicable to planets in
the HZ of K-dwarfs, although the respective estimates
should be lowered by 1-2 orders of magnitude.
4. CONCLUSIONS
We carried out numerical simulations of ancient (and
current) Mars to estimate the energy flux of SEPs
(specifically protons) impacting the planetary surface
and demonstrated that it is definitively lower than the
contributions from other well-known energy sources like
UV radiation and lightning. We also proved that energy
deposition of SEPs on the surface displays a distinctive
asymmetry with respect to both latitude and longitude.
We employed the energy flux of SEPs incident upon
the surface of ancient Mars to arrive at certain notewor-
thy conclusions. By drawing upon detailed experimen-
tal evidence, we hypothesized that a wide range of pre-
biotic molecules, such as RNA/DNA nucleobases, aro-
matic compounds and amino acids, could be synthesized
through irradiation by energetic protons. In particular,
the estimated maximum yields of proteinogenic amino
acids (∼ 107 kg/yr) and nucleobases (∼ 104 kg/yr) -
the building blocks of proteins and RNA/DNA respec-
tively - were potentially comparable to, or even orders
of magnitude higher than, some of the widely explored
pathways like electrical discharges and exogenous deliv-
ery by meteorites. We also suggested that SEPs may
have played an important indirect role in prebiotic syn-
thesis by initiating electrical discharges.
Subsequently, we generalized our analysis to encom-
pass exoplanets with Mars-like atmospheres around K-
and M-dwarfs. We conjectured that SEPs are perhaps
an important factor in synthesizing organic compounds
on exoplanets orbiting M-dwarfs, and that they might
mitigate the paucity of bioactive UV radiation (and UV-
mediated prebiotic chemistry). Hence, from the specific
viewpoint of prebiotic chemistry mediated by SEPs, it
is tempting to conclude that these exoplanets are more
conducive to abiogenesis.5
However, in light of the many uncertainties and mul-
tiple interlinked factors involved concerning the ori-
gin of life, all of the above statements should be re-
garded with due caution. Detailed experimental and
numerical follow-up studies based on the irradiation
of gaseous mixtures resembling Noachian and M-dwarf
(exo)planetary environments with MeV-GeV protons, in
conjunction with remote sensing and in situ searches for
organic molecules on Mars (Benner et al. 2000; ten Kate
2010; Vago et al. 2017),6 are necessary in order to fully
assess the veracity of our results. Nevertheless, it seems
plausible that SEPs might exert a profound influence
(positive and/or negative) on the origin and evolution of
life on habitable planets and moons (Heller et al. 2014),
both within and outside of our solar system.
The authors are very grateful to Stephen Bougher for
generously providing access to the M-GITM model, and
acknowledge valuable discussions with Andrew Knoll,
Janet Luhmann, Michael Summers, Lei Dai, Takuya
Shibayama and David Pawlowski. The detailed and
insightful report furnished by the referee is also much
appreciated. ML and AL were partly supported by
grants from the Breakthrough Prize Foundation for the
Starshot Initiative and Harvard University's Faculty of
Arts and Sciences, and by the Institute for Theory and
Computation (ITC) at Harvard University. CD was sup-
ported by the NASA Living With a Star Jack Eddy
Postdoctoral Fellowship Program, administered by the
University Corporation for Atmospheric Research. XF
and BMJ acknowledge support from NASA's MAVEN
mission. Resources for this work were provided by the
NASA High-End Computing (HEC) Program through
the NASA Advanced Supercomputing (NAS) Division
at Ames Research Center.
REFERENCES
5 Ironically, many of these characteristics are also possibly dele-
terious to the emergence and sustenance of complex surface-based
biospheres on Mars (Dartnell et al. 2007; Pavlov et al. 2012) and
Mars-like exoplanets in the HZ of M-dwarfs (Griessmeier et al.
2005; Tabataba-Vakili et al. 2016; Tilley et al. 2017).
6 In this context, it could be necessary to take into account
the oxidizing agents (e.g. perchlorate ions) incorporated into the
Martian regolith (Lasne et al. 2016), ostensibly resulting in an
oxidant extinction depth of . 5 m (Lammer et al. 2003).
Adam, Z. 2007, Astrobiology, 7, 852
Adcock, C. T., Hausrath, E. M., & Forster, P. M. 2013,
Nat. Geosci., 6, 824
Airapetian, V. S., Glocer, A., Gronoff, G., H´ebrard, E., &
Danchi, W. 2016, Nat. Geosci., 9, 452
Arndt, N. T., & Nisbet, E. G. 2012, Annu. Rev. Earth
Planet. Sci., 40, 521
Arvidson, R. E., Squyres, S. W., Bell, J. F., et al. 2014,
Science, 343, 1248097
11
Atri, D. 2017, Mon. Not. R. Astron. Soc. Lett., 465, L34
Attwater, J., Wochner, A., & Holliger, P. 2013, Nat. Chem.,
Cockell, C. S., & Knowland, J. 1999, Biol. Rev., 74, 311
Comisso, L., Lingam, M., Huang, Y.-M., & Bhattacharjee,
5, 1011
Bada, J. L. 2004, Earth Planet. Sci. Lett., 226, 1
-. 2013, Chem. Soc. Rev., 42, 2186
Baker, V. R. 2001, Nature, 412, 228
Ball, P., & Hallsworth, J. E. 2015, Phys. Chem. Chem.
A. 2016, Phys. Plasmas, 23, 100702
Cooper, J. F., Christian, E. R., Richardson, J. D., & Wang,
C. 2003, Earth Moon and Planets, 92, 261
Cronin, J. R., & Pizzarello, S. 1983, Adv. Space Res., 3, 5
Crosby, N. B., Aschwanden, M. J., & Dennis, B. R. 1993,
Phys., 17, 8297
Sol. Phys., 143, 275
Bartels-Rausch, T., Bergeron, V., Cartwright, J. H. E.,
Crutzen, P. J., Isaksen, I. S. A., & Reid, G. C. 1975,
et al. 2012, Rev. Mod. Phys., 84, 885
Science, 189, 457
Benner, S. A., Devine, K. G., Matveeva, L. N., & Powell,
Da Silva, L., Maurel, M.-C., & Deamer, D. 2015, J. Mol.
D. H. 2000, Proc. Natl. Acad. Sci. USA, 97, 2425
Evol., 80, 86
Benner, S. A., & Kim, H.-J. 2015, in Proc. SPIE, Vol. 9606,
Instruments, Methods, and Missions for Astrobiology
XVII, 96060C
Dartnell, L. R. 2011, Astrobiology, 11, 551
Dartnell, L. R., Desorgher, L., Ward, J. M., & Coates, A. J.
2007, Geophys. Res. Lett., 34, L02207
Benner, S. A., Kim, H.-J., & Carrigan, M. A. 2012, Acc.
Davis, W. L., & McKay, C. P. 1996, Orig. Life Evol.
Chem. Res., 45, 2025
Biosph., 26, 61
Benz, A. O. 2017, Living Rev. Sol. Phys., 14, 2
Berger, M. J., Coursey, J. S., Zucker, M. A., & Chang, J.
2011, Stopping-power and range tables for electrons,
protons, and helium ions, Tech. rep., NISTIR 4999,
ASTAR Program, NIST Physics Laboratory.
Deamer, D., & Weber, A. L. 2010, Cold Spring Harb.
Perspect. Biol., 2, a004929
Deamer, D. W. 1997, Microbiol. Mol. Biol. Rev., 61, 239
Deng, Y., Richmond, A. D., Ridley, A. J., & Liu, H.-L.
2008, Geophys. Res. Lett., 35, L01104
http://www.nist.gov/pml/data/star/index.cfm
Desai, M., & Giacalone, J. 2016, Living Rev. Sol. Phys., 13,
Boesswetter, A., Lammer, H., Kulikov, Y., Motschmann,
3
U., & Simon, S. 2010, Planet. Space Sci., 58, 2031
Borucki, W. J., & Chameides, W. L. 1984, Rev. Geophys.,
Di Achille, G., & Hynek, B. M. 2010, Nat. Geosci., 3, 459
Dobson, C. M., Ellison, G. B., Tuck, A. F., & Vaida, V.
22, 363
2000, Proc. Natl. Acad. Sci. USA, 97, 11864
Bougher, S. W., Bell, J. M., Murphy, J. R.,
Donaldson, D. J., Tervahattu, H., Tuck, A. F., & Vaida, V.
Lopez-Valverde, M. A., & Withers, P. G. 2006, Geophys.
Res. Lett., 33, L02203
2004, Orig. Life Evol. Biosph., 34, 57
Dong, C., Bougher, S. W., Ma, Y., et al. 2014, Geophys.
Bougher, S. W., Pawlowski, D., Bell, J. M., et al. 2015, J.
Res. Lett., 41, 2708
Geophys. Res. E, 120, 311
Dong, C., Huang, Z., Lingam, M., et al. 2017a, Astrophys.
Buccino, A. P., Lemarchand, G. A., & Mauas, P. J. D.
J. Lett., 847, L4
2007, Icarus, 192, 582
Dong, C., Jin, M., Lingam, M., et al. 2018, Proc. Natl.
Budin, I., & Szostak, J. W. 2010, Annu. Rev. Biophys., 39,
Acad. Sci. USA, 115, 260
245
Dong, C., Lingam, M., Ma, Y., & Cohen, O. 2017b,
Bywater, R. P., & Conde-Frieboesk, K. 2005, Astrobiology,
Astrophys. J. Lett., 837, L26
5, 568
Dong, C., Bougher, S. W., Ma, Y., et al. 2015a, J. Geophys.
Candelaresi, S., Hillier, A., Maehara, H., Brandenburg, A.,
Res. A, 120, 7857
& Shibata, K. 2014, Astrophys. J., 792, 67
Dong, C., Ma, Y., Bougher, S. W., et al. 2015b, Geophys.
Canfield, D. E., Glazer, A. N., & Falkowski, P. G. 2010,
Res. Lett., 42, 9103
Science, 330, 192
Carr, M. H., & Head, J. W. 2010, Earth Planet. Sci. Lett.,
Dwyer, J. R., & Uman, M. A. 2014, Phys. Rep., 534, 147
Ehlmann, B. L., & Edwards, C. S. 2014, Annu. Rev. Earth
294, 185
Planet. Sci., 42, 291
Chyba, C., & Sagan, C. 1992, Nature, 355, 125
Cleaves, H. J., & Miller, S. L. 1998, Proc. Natl. Acad. Sci.
Ehlmann, B. L., Mustard, J. F., Murchie, S. L., et al. 2011,
Nature, 479, 53
USA, 95, 7260
Ehrenfreund, P., Rasmussen, S., Cleaves, J., & Chen, L.
Cleaves II, H. J., Scott, A. M., Hill, F. C., et al. 2012,
2006, Astrobiology, 6, 490
Chem. Soc. Rev., 41, 5502
Ehrenfreund, P., Irvine, W., Becker, L., et al. 2002, Rep.
Cockell, C. S. 2014, Astrobiology, 14, 182
Prog. Phys., 65, 1427
12
Emslie, A. G., Dennis, B. R., Shih, A. Y., et al. 2012,
Hurowitz, J. A., & McLennan, S. M. 2007, Earth Planet.
Astrophys. J., 759, 71
Sci. Lett., 260, 432
Fair´en, A. G. 2010, Icarus, 208, 165
Fang, X., Bougher, S. W., Johnson, R. E., et al. 2013a,
Hurowitz, J. A., Grotzinger, J. P., Fischer, W. W., et al.
2017, Science, 356, eaah6849
Geophys. Res. Lett., 40, 1922
Jackman, C. H., Frederick, J. E., & Stolarski, R. S. 1980, J.
Fang, X., Liemohn, M. W., Kozyra, J. U., & Solomon, S. C.
Geophys. Res., 85, 7495
2004, J. Geophys. Res. A, 109, A04309
Jakosky, B. M., Nealson, K. H., Bakermans, C., Ley, R. E.,
Fang, X., Lummerzheim, D., & Jackman, C. H. 2013b, J.
& Mellon, M. T. 2003, Astrobiology, 3, 343
Geophys. Res. A, 118, 5369
Farquhar, J., Savarino, J., Jackson, T. L., & Thiemens,
Jakosky, B. M., & Phillips, R. J. 2001, Nature, 412, 237
Jakosky, B. M., & Shock, E. L. 1998, J. Geophys. Res., 103,
M. H. 2000, Nature, 404, 50
19359
Fassett, C. I., & Head, J. W. 2011, Icarus, 211, 1204
Ferris, J. P., & Hagan, W. J. 1984, Tetrahedron, 40, 1093
Ferris, J. P., Hill, A. R., Liu, R., & Orgel, L. E. 1996,
Nature, 381, 59
Feulner, G. 2012, Rev. Geophys., 50, RG2006
Feynman, J., Spitale, G., Wang, J., & Gabriel, S. 1993, J.
Geophys. Res., 98, 13281
Jakosky, B. M., Grebowsky, J. M., Luhmann, J. G., et al.
2015, Science, 350, 0210
Jolitz, R. D., Dong, C. F., Lee, C. O., et al. 2017, J.
Geophys. Res. A, 122, 5653
Joyce, G. F. 2002, Nature, 418, 214
Kasting, J. F. 2010, Nature, 464, 687
Kasting, J. F., Whitmire, D. P., & Reynolds, R. T. 1993,
Formisano, V., Atreya, S., Encrenaz, T., Ignatiev, N., &
Icarus, 101, 108
Giuranna, M. 2004, Science, 306, 1758
Kay, C., Opher, M., & Kornbleuth, M. 2016, Astrophys. J.,
Fox-Powell, M. G., Hallsworth, J. E., Cousins, C. R., &
826, 195
Cockell, C. S. 2016, Astrobiology, 16, 427
Knoll, A. H., Carr, M., Clark, B., et al. 2005, Earth Planet.
Gollihar, J., Levy, M., & Ellington, A. D. 2014, Science,
Sci. Lett., 240, 179
343, 259
Kobayashi, K., Kaneko, T., Saito, T., & Oshima, T. 1998,
Gopalswamy, N., Yashiro, S., & Akiyama, S. 2007, J.
Orig. Life Evol. Biosph., 28, 155
Geophys. Res. A, 112, A06112
Kobayashi, K., Kaneko, T., Tsuchiya, M., et al. 1995, Adv.
Griessmeier, J.-M., Stadelmann, A., Motschmann, U., et al.
Space Res., 15, 127
2005, Astrobiology, 5, 587
Kobayashi, K., Tsuchiya, M., Oshima, T., & Yanagawa, H.
Grott, M., Morschhauser, A., Breuer, D., & Hauber, E.
1990, Orig. Life Evol. Biosph., 20, 99
2011, Earth Planet. Sci. Lett., 308, 391
Kobayashi, K., Masuda, H., Ushio, K.-i., et al. 2001, Adv.
Gudel, M. 2007, Living Rev. Sol. Phys., 4, 3
Gurevich, A. V., Milikh, G. M., & Roussel-Dupre, R. 1992,
Space Res., 27, 207
Koonin, E. V., & Novozhilov, A. S. 2009, IUBMB life, 61,
Phys. Lett. A, 165, 463
99
Gurevich, A. V., Zybin, K. P., & Roussel-Dupre, R. A.
Kreysing, M., Keil, L., Lanzmich, S., & Braun, D. 2015,
1999, Phys. Lett. A, 254, 79
Nat. Chem., 7, 203
Halevy, I., & Head, III, J. W. 2014, Nat. Geosci., 7, 865
Halevy, I., Zuber, M. T., & Schrag, D. P. 2007, Science,
318, 1903
Hannah, I. G., Hudson, H. S., Battaglia, M., et al. 2011,
Kvenvolden, K., Lawless, J., Pering, K., et al. 1970, Nature,
228, 923
Lambert, J.-F. 2008, Orig. Life Evol. Biosph., 38, 211
Lammer, H., Lichtenegger, H. I. M., Kolb, C., et al. 2003,
Space Sci. Rev., 159, 263
Icarus, 165, 9
Hassler, D. M., Zeitlin, C., Wimmer-Schweingruber, R. F.,
Lasne, J., Noblet, A., Szopa, C., et al. 2016, Astrobiology,
et al. 2014, Science, 343, 1244797
16, 977
Hazen, R. M., & Sholl, D. S. 2003, Nat. Mater., 2, 367
Hazen, R. M., & Sverjensky, D. A. 2010, Cold Spring Harb.
Lillis, R. J., Frey, H. V., & Manga, M. 2008, Geophys. Res.
Lett., 35, L14203
Perspect. Biol., 2, a002162
Lillis, R. J., Robbins, S., Manga, M., Halekas, J. S., & Frey,
Heller, R., Williams, D., Kipping, D., et al. 2014,
H. V. 2013, J. Geophys. Res. E, 118, 1488
Astrobiology, 14, 798
Higgs, P. G., & Pudritz, R. E. 2009, Astrobiology, 9, 483
Hodos´an, G., Helling, C., Asensio-Torres, R., Vorgul, I., &
Lingam, M., & Loeb, A. 2017a, Astrophys. J., 848, 41
-. 2017b, submitted to J. Cosmol. Astropart. Phys.,
arXiv:1710.11134
Rimmer, P. B. 2016, Mon. Not. R. Astron. Soc., 461, 3927
-. 2017c, submitted to Int. J. Astrobiol., arXiv:1711.09908
13
-. 2017d, Astrobiology, arXiv:1707.04594
-. 2017e, Int. J. Astrobiol., arXiv:1707.02996
-. 2017f, Astrophys. J. Lett., 846, L21
L´opez-Puertas, M., Funke, B., Gil-L´opez, S., et al. 2005, J.
Geophys. Res. A, 110, A09S43
Luisi, P. L. 2016, The Emergence of Life: From Chemical
Origins to Synthetic Biology (Cambridge Univ. Press)
Lummerzheim, D., Rees, M. H., & Anderson, H. R. 1989,
Planet. Space Sci., 37, 109
Ma, Y. J., Russell, C. T., Fang, X., et al. 2017, J. Geophys.
Res. A, 122, 1714
Orgel, L. E. 2004, Crit. Rev. Biochem. Mol. Biol., 39, 99
Owen, T. 1992, in Mars, ed. H. H. Kieffer, B. M. Jakosky,
C. W. Snyder, & M. S. Matthews (Univ. of Arizona
Press), 818–834
Pascal, R. 2012, J. Sys. Chem., 3, 3
Patel, B. H., Percivalle, C., Ritson, D. J., Duffy, C. D., &
Sutherland, J. D. 2015, Nat. Chem., 7, 301
Pavlov, A. A., Vasilyev, G., Ostryakov, V. M., Pavlov,
A. K., & Mahaffy, P. 2012, Geophys. Res. Lett., 39,
L13202
Pearce, B. K. D., & Pudritz, R. E. 2016, Astrobiology, 16,
Maehara, H., Shibayama, T., Notsu, Y., et al. 2015, Earth,
853
Planets, and Space, 67, 59
Pizzarello, S., & Shock, E. 2017, Orig. Life Evol. Biosph.,
Maehara, H., Shibayama, T., Notsu, S., et al. 2012, Nature,
47, 249
485, 478
Maher, K. A., & Stevenson, D. J. 1988, Nature, 331, 612
Martin, W., Baross, J., Kelley, D., & Russell, M. J. 2008,
Nat. Rev. Microbiol., 6, 805
McCollom, T. M. 2013, Annu. Rev. Earth Planet. Sci., 41,
207
McKay, C. P. 1997, Orig. Life Evol. Biosph., 27, 263
-. 2010, Cold Spring Harb. Perspect. Biol., 2, a003509
Mekhaldi, F., Muscheler, R., Adolphi, F., et al. 2015, Nat.
Commun., 6, 8611
Melott, A. L., & Thomas, B. C. 2011, Astrobiology, 11, 343
Mewaldt, R. A. 2006, Space Sci. Rev., 124, 303
Mewaldt, R. A., Looper, M. D., Cohen, C. M. S., et al.
Price, P. B. 2007, FEMS Microbiol. Ecol., 59, 217
Priest, E. 2014, Magnetohydrodynamics of the Sun
(Cambridge Univ. Press)
Ranjan, S., & Sasselov, D. D. 2017, Astrobiology, 17, 169
Ranjan, S., Wordsworth, R., & Sasselov, D. D. 2017a,
Astrophys. J., 843, 110
-. 2017b, Astrobiology, 17, 687
Rapf, R. J., & Vaida, V. 2016, Phys. Chem. Chem. Phys.,
18, 20067
Reames, D. V. 2013, Space Sci. Rev., 175, 53
Ribas, I., Guinan, E. F., Gudel, M., & Audard, M. 2005,
Astrophys. J., 622, 680
Ridley, A. J., Deng, Y., & T´oth, G. 2006, J. Atmos.
2012, Space Sci. Rev., 171, 97
Sol.-Terr. Phys., 68, 839
Mileikowsky, C., Cucinotta, F. A., Wilson, J. W., et al.
2000, Icarus, 145, 391
Ross, D. S., & Deamer, D. 2016, Life, 6, 28
Rugheimer, S., Segura, A., Kaltenegger, L., & Sasselov, D.
Miller, S. L. 1953, Science, 117, 528
-. 1998, in The Molecular Origins of Life, ed. A. Brack
2015, Astrophys. J., 806, 137
Ruiz-Mirazo, K., Briones, C., & de la Escosura, A. 2014,
(Cambridge Univ. Press), 59–85
Chem. Rev., 114, 285
Miller, S. L., & Urey, H. C. 1959, Science, 130, 245
Miyakawa, S., Cleaves, H. J., & Miller, S. L. 2002a, Orig.
Russell, M. J., Barge, L. M., Bhartia, R., et al. 2014,
Astrobiology, 14, 308
Life Evol. Biosph., 32, 209
Miyakawa, S., Yamanashi, H., Kobayashi, K., Cleaves,
H. J., & Miller, S. L. 2002b, Proc. Natl. Acad. Sci. USA,
99, 14628
Sagan, C., & Mullen, G. 1972, Science, 177, 52
Sanuki, T., Motoki, M., Matsumoto, H., et al. 2000,
Astrophys. J., 545, 1135
Schrijver, C. J., Beer, J., Baltensperger, U., et al. 2012, J.
Monnard, P.-A., & Szostak, J. W. 2008, J. Inorg. Biochem.,
Geophys. Res. A, 117, A08103
102, 1104
Schubert, G., Russell, C. T., & Moore, W. B. 2000, Nature,
Morowitz, H., & Smith, E. 2007, Complexity, 13, 51
Mutschler, H., Wochner, A., & Holliger, P. 2015, Nat.
408, 666
Schulze-Makuch, D., Fair´en, A. G., & Davila, A. F. 2008,
Chem., 7, 502
Int. J. Astrobiol., 7, 117
Nava-Sedeno, J. M., Ortiz-Cervantes, A., Segura, A., &
Domagal-Goldman, S. D. 2016, Astrobiology, 16, 744
Newman, M. E. J. 2005, Contemp. Phys., 46, 323
Nisbet, E. G., & Sleep, N. H. 2001, Nature, 409, 1083
Odert, P., Leitzinger, M., Hanslmeier, A., & Lammer, H.
Scott, C. J., Harrison, R. G., Owens, M. J., Lockwood, M.,
& Barnard, L. 2014, Environ. Res. Lett., 9, 055004
Segura, A., & Navarro-Gonz´alez, R. 2005, Geophys. Res.
Lett., 32, L05203
Shibata, K., Isobe, H., Hillier, A., et al. 2013, Publ. Astron.
2017, Mon. Not. R. Astron. Soc., 472, 876
Soc. Jpn, 65, 49
14
Shibayama, T., Maehara, H., Notsu, S., et al. 2013,
Tilley, M. A., Segura, A., Meadows, V. S., Hawley, S., &
Astrophys. J. Suppl., 209, 5
Shields, A. L., Ballard, S., & Johnson, J. A. 2016, Phys.
Davenport, J. 2017, submitted to Astrobiology,
arXiv:1711.08484
Rep., 663, 1
Tosca, N. J., & Knoll, A. H. 2009, Earth Planet. Sci. Lett.,
Siingh, D., Singh, R. P., Singh, A. K., et al. 2011, Surveys
286, 379
in Geophysics, 32, 659
Tosca, N. J., Knoll, A. H., & McLennan, S. M. 2008,
Smith, D. S., Scalo, J., & Wheeler, J. C. 2004, Orig. Life
Science, 320, 1204
Evol. Biosph., 34, 513
Solomon, S. 1999, Rev. Geophys., 37, 275
Solomon, S. C., Aharonson, O., Aurnou, J. M., et al. 2005,
Science, 307, 1214
Spiegel, D. S., & Turner, E. L. 2012, Proc. Natl. Acad. Sci.
USA, 109, 395
Squyres, S. W., & Knoll, A. H. 2005, Earth. Planet. Sci.
Lett., 240, 1
Trinks, H., Schroder, W., & Biebricher, C. K. 2005, Orig.
Life Evol. Biosph., 35, 429
Tuck, A. 2002, Surv. Geophys, 23, 379
Vago, J. L., Westall, F., Pasteur Instrument Team, et al.
2017, Astrobiology, 17, 471
Walker, S. I. 2017, Rep. Prog. Phys., 80, 092601
Webb, D. F., & Howard, T. A. 2012, Living Rev. Sol.
Stribling, R., & Miller, S. L. 1987, Orig. Life Evol. Biosph.,
Phys., 9, 3
17, 261
Stueken, E. E., Anderson, R. E., Bowman, J. S., et al. 2013,
Weber, A. L., & Miller, S. L. 1981, J. Mol. Evol., 17, 273
Webster, C. R., Mahaffy, P. R., Atreya, S. K., et al. 2015,
Geobiology, 11, 101
Science, 347, 415
Summers, M. E., Lieb, B. J., Chapman, E., & Yung, Y. L.
Westall, F., Loizeau, D., Foucher, F., et al. 2013,
2002, Geophys. Res. Lett., 29, 24.1
Astrobiology, 13, 887
Sutherland, J. D. 2016, Angew. Chem. Int. Ed., 55, 104
-. 2017, Nat. Rev. Chem., 1, 0012
Tabataba-Vakili, F., Grenfell, J. L., Griessmeier, J.-M., &
Rauer, H. 2016, Astron. Astrophys., 585, A96
Takahashi, T., Mizuno, Y., & Shibata, K. 2016, Astrophys.
J. Lett., 833, L8
ten Kate, I. L. 2010, Astrobiology, 10, 589
Tian, F., Kasting, J. F., & Solomon, S. C. 2009, Geophys.
Res. Lett., 36, L02205
Tian, F., Kasting, J. F., & Zahnle, K. 2011, Earth Planet.
Sci. Lett., 308, 417
Wordsworth, R. D. 2016, Annu. Rev. Earth Planet. Sci., 44,
381
Yashiro, S., Gopalswamy, N., Michalek, G., et al. 2004, J.
Geophys. Res. A, 109, A07105
Youngblood, A., France, K., Parke Loyd, R. O., et al. 2017,
Astrophys. J., 843, 31
Zahnle, K., Schaefer, L., & Fegley, B. 2010, Cold Spring
Harb. Perspect. Biol., 2, a004895
Zaia, D. A. M., Zaia, C. T. B. V., & de Santana, H. 2008,
Orig. Life Evol. Biosph., 38, 469
|
1901.02383 | 1 | 1901 | 2019-01-08T16:03:35 | The secondary transit of the hot Jupiter WASP-121b at 2 $\mu$m | [
"astro-ph.EP"
] | Ground-based observations of the secondary transit in the 2MASS K band are presented for the hot Jupiter WASP-121b. These are the first occultation observations of an extrasolar planet carried out with an instrument attached to a 1m-class telescope (SMARTS' 1.3 m). We find a highly significant transit depth of (0.228 +/- 0.023)%. Together with the Hubble Space Telescope near infrared emission spectrum, current data support more involved atmosphere models with species producing emission/absorption features, rather than simple smooth black body emission. Analysis of the time difference between the primary and secondary transits and the durations of these events yield an eccentricity of e=0.0207 +/- 0.0153, which is consistent with the earlier estimates of low/zero eccentricity, but with a smaller error. Together with the existing K-band data on other systems, WASP-121b lends further support to the lack of efficient heat transport between the day and night sides for nearly all Hot Jupiters. | astro-ph.EP | astro-ph | Astronomy& Astrophysics manuscript no. wasp121v3aph
January 9, 2019
c(cid:13) ESO 2019
9
1
0
2
n
a
J
8
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
3
8
3
2
0
.
1
0
9
1
:
v
i
X
r
a
The secondary transit of the hot Jupiter WASP-121b at 2 µm
G´eza Kov´acs1 and Tam´as Kov´acs2
1 Konkoly Observatory of the Hungarian Academy of Sciences, Budapest, 1121 Konkoly Thege ut. 13-15, Hungary
e-mail: [email protected]
2 Institute of Theoretical Physics, Eotvos University, Budapest, 1117 P´azm´any P´eter s´et´any 1A
Received September ?? 2018 / Accepted ?? 2018
ABSTRACT
Ground-based observations of the secondary transit in the 2MASS K band are presented for the hot Jupiter WASP-121b. These are
the first occultation observations of an extrasolar planet carried out with an instrument attached to a 1 m-class telescope (SMARTS'
1.3 m). We find a highly significant transit depth of (0.228±0.023)%. Together with the Hubble Space Telescope near infrared emission
spectrum, current data support more involved atmosphere models with species producing emission/absorption features, rather than
simple smooth black body emission. Analysis of the time difference between the primary and secondary transits and the durations
of these events yield an eccentricity of e = 0.0207 ± 0.0153, which is consistent with the earlier estimates of low/zero eccentricity,
but with a smaller error. Together with the existing K-band data on other systems, WASP-121b lends further support to the lack of
efficient heat transport between the day and night sides for nearly all Hot Jupiters.
Key words. Planets and satellites: atmospheres -- Methods: data analysis
1. Introduction
When combined with other pieces of information (such as planet
mass), the low thermal radiation of extrasolar planets is a di-
rect evidence of their substellar nature. In addition to this inde-
pendent verification, measuring the radiation spectrum yields a
wealth of information on atmospheric structure and basic orbital
parameters. Due to their low temperatures (relative to their host
stars), the best chance of detection is obviously in the infrared.
From the first detection, employing the mid-infrared instrument
of the Spitzer Space Telescope by Deming et al. (2005), many
systems have been observed not only by space-based, but also by
ground-based instruments attached to 4-meter class telescopes.
Here we report the multiple detection of the secondary transit1
of the Hot Jupiter (HJ) WASP-121b in the 2MASS K band by
A Novel Dual Imaging CAMera (ANDICAM) attached to the
1.3 m telescope of the SMARTS Consortium2.
The transiting extrasolar planetary system WASP-121 was
discovered by Delrez et al. (2016) using the wide-field tele-
scopes of the SuperWASP project (Pollacco et al. 2006, see
also Anderson et al. 2018 for the latest update). The analysis of
these and the subsequent spectroscopic followup observations
revealed that WASP-121b is a very hot Jupiter, with a maxi-
mum photospheric temperature above 3000 K. This is not sur-
prising, since the planet orbits an F star rather close, with a pe-
riod of 1.27 days. The close orbit, the extended planet's radius3
1 Throughout this papers, we use also the word 'occultation' for the
event of secondary transit.
2 For additional details on the instrument, access images, raw and
systematics corrected light curves, please visit
http://www.astro.yale.edu/smarts/1.3m.html
http://archive.noao.edu/search/query/
http://cdsweb.u-strasbg.fr/ .
of Rp = 1.87 RJ with standard mass of Mp = 1.18 MJ imply
rather strong tidal dissipation, leading to Roche-lobe filling and
then to a speedy disruption within some hundred million years,
assuming a stellar tidal dissipation factor Qstar < 108 (see Delrez
et al. 2016). In addition, the planet has probably experienced a
strong dynamical interaction with some nearby third body, as can
be inferred from the large projected spin-orbit angle of 258◦ ± 5◦
(see Delrez et al. 2016). Interestingly, secondary transits were
observed multiple times by the same authors in the Sloan-z' band
by the 60 cm TRAPPIST telescope -- to our knowledge, this is
the first occultation detection from the ground by a telescope
of this size. The significant detection of the occultation depth
of a mere (0.060 ± 0.013)% resulted in the first direct estima-
tion of the planet temperature. Important followup observations
(both during the primary and secondary transits) have been made
by Evans et al. (2016, 2017, 2018) by using the Hubble Space
Telescope's Wide Field Camera 3 in the near infrared, Spitzer's
IRAC detector at 3.6 µm and HST/STIS in the UV. These data
indicate a weak H2O emission during occultation and absorp-
tion during transit, implying temperature inversion due to some
high-altitude absorber. In spite of the successful fit of the HST
emission spectrum, and quite currently the transmission spec-
trum observed by the same instrument, the authors caution for
the non-uniqueness of the solution (e.g., type of absorber, pre-
cision of the fit at different wavelengths, etc.). These issues are
not unique to WASP-121b, they are also present in other very
hot Jupiters (e.g., WASP-33b, Kepler-13Ab -- see Parmentier et
al. 2018). In spite of the considerable progress made in the past
ten years, there is a substantial lack of understanding the rela-
tions between the physical parameters of the systems and the
thermal properties of their planets (see the uniform analysis by
Adams & Laughlin 2018 of 10 systems with full infrared phase
curves).
3 The radius quoted here is the one that appears in the Abstract of the
discovery paper. We use the radius based on the HST measurements of
Evans et al. (2017) -- see Sect. 4.
The purpose of this work is to add a flux value to the emis-
sion spectrum of WASP-121b at a single waveband and thereby
increase the number of constraints on future atmosphere mod-
1
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
eling of the planet. Furthermore, timing estimates are presented
to give more stringent limits on the orbital eccentricity, a valu-
able parameter for the analysis of the dynamical history of the
system.
2. Observations and the method of analysis
Photometric observations in the 2MASS K and Cousins I bands
(effective wavelengths of 2.2 µm and 0.8 µm, respectively) have
been made by using the ANDICAM instrument in a beam-
splitting mode, allowing simultaneous data acquisition in the two
bands.4 On each night, the target was monitored continuously,
by allowing ample amount of pre- and after-event time (permit-
ted by the actual sky position) to reliably fix the out-of-transit
(OOT) baseline for the event lasting almost for 3 hours. An ex-
posure time of 15 -- 20 s was used, resulting in an overall cadence
of 70 -- 500 s, due to the overheads, related to read-outs, vary-
ing movements due to dithering and other data acquisition steps.
The observing log with some associated parameters is given in
Table 1.
Table 1. List of secondary transit observations of WASP-121b
in the near infrared.
Date [UT]
02/26/2016
01/11/2017
01/25/2017
Tocc [HJD]
Ttot [hours]
2457444.64855
2457764.65485
2457778.67903
4.86
6.13
6.64
Ntot
177
203
205
Notes:
Tocc stands for the time of the center of occultation as estimated from
the ephemeris given by Delrez et al. (2016) for the primary transit in
their Table 4. Ttot and Ntot are the total observing time and data points
gathered.
For the K-band observations dithering was used to decrease
the higher sensitivity against detector non-uniformity in the near
infrared. We found this method useful, as we did not have a pri-
ori information on pixel sensitivity. Admittedly, this method has
also some risk, since by testing different parts of the CCD, we
may bump into bad positions, leading to light curves of larger
scatter associated with the particular dither position. All in all,
we think that our strategy has proven to be useful and led to a
higher quality result at the end.
By stacking several images, we show the dither pattern in
Fig. 1 for one of the nights. The number of dither positions
changed from night to night, and their durations were also not
the same. The image (that has already been corrected for flat
field) spectacularly exhibits sequences of rings, reminiscent of
the trace of some earlier drops of dew. In spite of their high vis-
ibility, their effect has been proven to be less damaging for the
data quality than the varying pixel sensitivity (that is consider-
able more difficult to spot, because they lack the type of spatial
correlation the rings have).
In producing the photometric time series to be used in the
derivation of the basic occultation parameters, we proceed as
follows. First we compute simple relative fluxes at various, but
fixed circular apertures from 10 to 20 pixel radii with an incre-
ment of 2 pixels. After a lot of experimenting, and inspecting the
4 Unfortunately, the signal -- hampered by weather and instrumental
limitations -- in the Cousins I band was too weak to yield any useful
planet atmospheric constraint, so we decided not to deal with it.
2
WASP-121 (Feb 26, 2016)
500
400
n
o
i
t
i
s
o
p
l
e
x
p
i
300
200
100
0
0
100
200
pixel position
300
400
500
95000
90000
85000
s
t
n
u
o
c
80000
75000
Fig. 1. Dithering pattern used during the near infrared observa-
tions. The image shows the 2.4′ × 2.4′ FOV of the ANDICAM
near infrared camera, attached to the 1.3 m telescope of
the SMARTS Consortium. The target (WASP-121=2MASS
07102406-3905506) is in the middle,
the comparison star
2MASS 07102364-3905561 is in the lower left corner. North is
on the top West is on the left. Circles around the target and the
comparison star show the aperture sizes used to estimate the star
and background fluxes.
#6
#6
X
U
L
F
1.04
1.03
1.02
1.01
1.00
0.99
0.98
0.97
#13
#9
0.40
0.45
0.50
0.55
PHASE
0.60
0.65
Fig. 2. Simple photometric flux ratios ordered by the orbital pe-
riod (pale dots). Some dithers are annotated to see the nightly
trends (or the lack of them, i.e., #9). Dither #6 (gray dots) is
plotted also after employing zero point shift and detrending by
the position vector (black dots, see Sect. 2 for details).
final product of the full detrending procedure to be described be-
low, we find that the aperture with the pixel radius of 16 yields a
light curve with the smallest scatter. All results presented in this
paper refer to the above aperture size.
By using the relative fluxes (target over comparison star flux,
hereafter raw flux) and folding the data with the orbital period,
we can examine if we can see some sign of an occultation event.
The result is shown in Fig. 2. The pale dots show that the raw
fluxes are very noisy, and the event with the expected depth of
0.1 -- 0.2% is hopelessly buried in the noise. We can find out the
reason of this somewhat unexpected high level of noise by ex-
amining the individual light curves (LCs) associated with the
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
various dither positions5. Indeed, we see from the highlighted
LCs that there is a strong dependence on the dither position,
leading to both zero point shifts and nightly trends. Therefore,
(not entirely unexpectedly), we must employ some detrending
method that is likely the cause of the trends and zero point shifts.
The detrending step is vital and therefore quite common in the
extraction of planetary signals in general, and, in particular, in
the derivation of wavelength-dependent transit depths for the
exquisite accuracy needed to estimate emission or transmission
spectra (e.g., Stevenson et al. 2012; Kreidberg et al. 2015).
It is well-known that ground-based instruments detect stel-
lar light deformed by the multiplicative noise and systematics
originating from the Earth's atmosphere and from the environ-
ment/instrument. In addition, we also have an additive noise
source, coming from the sky background
F = F0 × Tatm × Tenv + Fbg ,
(1)
where F is the detected, F0 is the true stellar flux. The trans-
mission functions of the Earth's atmosphere and the instrument
are denoted by Tatm and Tenv, respectively, whereas the back-
ground flux by Fbg. In traditional photometric reductions the at-
mospheric and instrumental effects are filtered out with the aid
of comparison stars near the target, by using the assumption of
the close similarity in the transmission functions for the target
and its neighboring companions. However, when higher accu-
racy is required, this method usually fails, because of the lack of
complete equivalence between the transmission functions of the
target and the comparison stars (and, for faint targets, there is
also the issue of the presence of the additive background noise).
Due to the lack of obvious exact solution of the problem
(similarly to the methodology followed in other studies, e.g.,
Bakos et al. 2010, Delrez et al. 2016), we opt to an approximate
one. Here we take the logarithm of the target to comparison star
flux ratios F/F c, and fit the data with the linear combination of
the presumed signal and certain external photometric parame-
ters (e.g., position, PSF width). In addition, we treat each LC
of the different dither positions individually, with particular zero
points and trends (but with the same underlying signal). That is,
we Least Square minimize the following expression
D =
M
Nj
Xj=1
Xi=1
wj" log
Fj(i)
F c
j (i)
− Ej(i)#2
,
Ej(i) = a0,j + ax,jXj(i) + ay,jYj(i) + A log(Ftrap(i)) .
(2)
(3)
Here all data are sorted by the orbital phase. We assume that
there are M dither positions altogether with Nj data points at the
j-th dither. Since our extensive tests showed that neither arbi-
trary polynomial nor additional external parameters are needed
to reach a high signal-to-noise ratio (S /N) detection, we use
only the pixel position components (X, Y) of the target to cor-
rect for instrumental effects. The stellar flux during the occulta-
tion is approximated by a trapezoidal function Ftrap with fixed
ingress/egress time, duration and transit center of 0.015, 0.120
and 2457764.65485 days, respectively, corresponding to those
given by Delrez et al. (2016). The weights {w} are constant
for the same dither index and proportional to the reciprocal of
the variance of the residuals around the best-fitting trapezoidal.
Since the solution is not known, the weights are iterated during
the process of solution. At the end, the data are converted back
to relative intensities, with an OOT normalization of 1.0 for the
5 Please note that all dither positions are counted and their indices
increase toward more recent nights of observation.
fitted trapezoidal. The error of the occultation depth is computed
as follows
σ(δocc) =
1 −
Np
N
−1/2
s14
N2
14
+
soot
N2
oot
1/2
,
(4)
where s14 and soot are sums of the squared residuals in the in- and
out-of-transit phases, respectively. Akin to these are the number
of data points N14 and Noot. The factor in front (with Np number
of the parameters fitted to N total number of data points) is for
the debiasing of the error due to the decrease of the degrees of
freedom, because of parameter fitting. The S /N of the detection
is the ratio of the average transit depth to the above error
S /N =
1
σ(δocc)
1
N14
N14
Xi=1
Ftrap(i) .
3. Occultation parameters
(5)
First we fix all secondary transit parameters (but the occulta-
tion depth) by assuming circular orbit and the validity of the pa-
rameters derived for the primary transit by Delrez et al. (2016).
Following the procedure described in Sect. 2, we compute the
best fitting occultation depth under various conditions, concern-
ing the number of points clipped and the dither LCs omitted.
The result is shown in Table. 2. Except perhaps for the extreme
choices of data trimming parameters (Ncut
σ ), the occul-
tation depth is relative stable. To avoid too sparsely populated
dither LCs, and not to 'overtrim' the data, we opt for the case
of Ncut
σ = 4. The folded LC obtained in this
dith
way is shown in Fig. 3. The resulting secondary transit depth is
0.00228 ± 0.00023.
= 10 and Ncut
dith and Ncut
Table 2. Occultation depths for WASP-121 in the 2MASS K
band
Ncut
dith
0
0
0
0
10
10
10
10
15
15
15
15
Ncut
σ
∞
5
4
3
∞
5
4
3
∞
5
4
3
Npar
70 + 0
70 + 8
70 + 11
70 + 25
61 + 0
61 + 8
61 + 11
61 + 25
52 + 0
52 + 8
52 + 11
52 + 25
N
585
585
585
585
561
561
561
561
520
520
520
520
S /N
8.1
8.4
8.8
9.2
8.2
8.5
8.9
9.3
8.2
8.6
8.9
9.3
σfit
0.00272
0.00264
0.00264
0.00256
0.00275
0.00266
0.00266
0.00257
0.00281
0.00270
0.00270
0.00260
δocc
0.00212
0.00215
0.00224
0.00227
0.00217
0.00220
0.00228
0.00231
0.00231
0.00233
0.00243
0.00244
dith is the lower limit on the number of data points per dither position.
σ is the number of standard deviations used in the clipping of the
Notes:
Ncut
Ncut
data points. Npar is the number of parameters fitted, plus the number of
data points omitted (Np in Eq. 4 includes both of these). Items in the
last two columns (unbiased estimates of the standard deviation of the
residuals and occultation depth) refer to the OOT= 1 normalization as
described in Sect. 2.
To check the level of the systematics filtering, we compute
the autocorrelation function (ACF) of the residuals after sub-
tracting the best-fitting trapezoidal as shown in Fig. 3. In the
units of the orbital period, the ACF is computed with steps of
0.00123 up to 0.115, i.e. close to the length of the full transit
3
1.01
X
U
L
F
1.00
0.99
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
SMARTS
2MASS K
0.02
0.01
c
T
∆
0
-0.01
-0.02
0.40
0.45
0.50
0.55
PHASE
0.60
0.65
0.1
0.11
0.13
0.14
0.12
t14
7.6e-06
7.5e-06
7.4e-06
7.3e-06
7.2e-06
7.1e-06
7e-06
6.9e-06
Fig. 3. Systematics-filtered folded flux ratios normalized to 1.0
in the out-of-transit part. Average fluxes (in 30 phase bins) are
shown by blue dashes, the best fitting trapezoidal secondary tran-
sit approximation is plotted by yellow continuous line.
Fig. 5. Intensity plot for the unbiased estimate of the variance of
the residuals between the data and the occultation model scanned
in the parameter space of the displacement of the occultation
center ∆Tc and the duration of the event t14. We employ iter-
ative 4σ clipping to find the best solution for each parameter
combination.
0.8
0.4
0.0
0.000
0.005
0.010
n
o
i
t
a
l
e
r
r
o
c
d
e
z
i
l
a
m
r
o
N
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
0.00
0.02
0.04
0.06
0.08
0.10
0.12
Shift [phase]
Fig. 4. Blue dots: autocorrelation function (ACF) of the resid-
uals of the trapezoidal fit to the final dataset shown in Fig. 3.
Red dots: ACF of generated uncorrelated noise. Error bars are
for the standard deviations of the the ACF values of the random
datasets. The time lag is given in the units of the orbital period.
The inset shows the immediate neighborhood of ACF at zero
time shift (given in the units of the orbital period).
TRAPPIST
Sloan z'
0.02
0.01
c
T
∆
0
-0.01
-0.02
0.1
0.11
0.13
0.14
0.12
t14
1.024e-05
1.023e-05
1.022e-05
1.021e-05
1.02e-05
1.019e-05
1.018e-05
1.017e-05
Fig. 6. As in Fig. 5, but for the TRAPPIST data. The better con-
trast (in spite of the larger residual scatter) of the best solution is
likely accounted for by the more than ten times larger number of
data points than for the SMARTS observations.
event. As a sanity check, we also compute the ACF for many
Gaussian white noise realizations. The result is shown in Fig. 4.
We see that the residuals are almost uncorrelated. The basic cor-
relation length is under ∼ 0.005 in the units of the orbital period.
This value is less than one half of the ingress duration. It seems
that the de-correlation method applied yields nearly white noise
residuals, supporting the validity of the pure statistical error esti-
mation given by Eq. 4. We also note that similar short-time-scale
correlations are observable in other studies dealing with system-
atics, in particular in the analysis of the HST data by Evans et
al. (2017).
Although the noise is rather high, the relative large number
of data points led to a high-S /N detection. Therefore, it is tempt-
ing to see if our assumption on the applicability of the primary
transit parameters is really held, and if there is a chance to fur-
ther constrain the eccentricity by the best-fitting occultation cen-
ter and event duration. To this aim we map the quality of the fit
as a function of ∆Tc (tested occultation center time minus the
one calculated from the primary transit with the assumption of
circular orbit) and t14 (occultation duration).
In addition to our data, to examine further the issue of ec-
centricity, the secondary transit data of Delrez et al. (2016) are
also investigated. Since the observations were made in the Sloan
z' band, the signal is considerably shallower than in the 2MASS
K (Ks) band. Nevertheless, the number of data points (6260 flux
measurements on seven nights) compensates for this, and yields
a confident detection of S /N = 7.5, with δocc = 0.000697 ±
0.000081 and a residual standard deviation of 0.003190. This
depth is larger by 0.000096 than the one derived by Delrez et
al. (2016), but the difference is within 1σ, and could be ac-
counted for by the lower number of detrending parameters used
in our code. We found it satisfactory to use only the pixel coordi-
nates, and avoid to correct with a polynomial and other parame-
ters, since these do not yield an appreciable improvement in the
quality of the fit, and, in addition, may lead to a depression of
the occultation depth.
4
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
tricity
The (∆Tc, t14) maps are shown in Figs. 5 and 6 for the Ks and
Sloan z' data, respectively. As expected, the topology of both
maps confirms the rather small (if any) deviations from the pa-
rameters predicted by the primary transit with the assumption of
circular orbit. Furthermore, the Sloan z' data are more restric-
tive than the Ks data, even though the S /N value is higher for the
latter. This is because the parameter maps yield information also
on the sensitivity of the solution on the neighboring parameter
values and not only on a specific combination of the parameters,
that might be better or worse, depending on the functional form
of the variance on these parameters and noise level. The better
quality of the Sloan z' data is also visible from the nearly three
times smaller error of the derived occultation depth.
Table 3. Observed secondary transit times for WASP-121b
Dataset
SMARTS (K)
TRAPPIST (z')
HST ('white')
Spitzer (3.6)
Tocc [BJD] O − C [d]
−0.0080
±0.0023
−0.0040
±0.0027
0.0004
±0.0004
−0.0013
±0.0007
2457764.6469
±0.0023
2456762.5594
±0.0027
2457703.4588
±0.0004
2457783.7774
±0.0007
t14
Source
0.116 KOV18
±0.005
0.128 KOV18
±0.007
−
EVA17
−
EVA17
Notes:
EVA17: Evans et al. (2017) -- KOV18: this paper (the source of the
TRAPPIST data is Delrez et al. 2016) -- The O-C values are computed
in respect of the ephemerides predicted from the primary transit as
given by Delrez et al. (2016), assuming circular orbit. -- See text for the
equality of the errors on Tocc and O − C. -- Evans et al. (2017) do not
supply occultation duration values.
The currently available secondary transit parameters are
summarized in Table 3. In the case of the KOV18 items the errors
have been computed in the following way. Once the best-fitting
trapezoidal was found, we added Gaussian white noise with the
observed standard deviations of the residuals corresponding to
this solution, and then the best-fitting trapezoidal to these simu-
lated data was searched for. By repeating the process 500 times
we arrived to statistically stable estimates of the formal errors.
The ingress/egress time was always fixed to the observed val-
ues given by the primary transit data of Delrez et al. (2016), and
we did the same also with the remaining parameters, depending
which parameter was tested for errors (e.g., in the case of the
occultation center, we fixed the duration and the ingress/egress
times). Although this approach is primarily dictated by keep-
ing the execution time within a reasonable limit, our error esti-
mates for the moment of the occultation time is in perfect agree-
ment with the one predicted by the analytic formula of Deeg &
Tingley (2017). The errors of O − C are taken equal to those of
Tocc, because the errors of the computed occultation times (C)
have been proven to be negligible.
We see that the available observations suggest a small (or
zero) eccentricity. Since the more precise estimation requires
also the knowledge of the transit duration, the lack of this param-
eter for the most accurate HST and Spitzer observations makes
us unable to include these data in the analysis. Therefore, we use
only the occultation parameters derived from the SMARTS and
TRAPPIST observations.
2
π
2
e =
∆Tc
P
2
+
r14 − 1
r14 + 1
1
2
,
(6)
where P is the orbital period, ∆Tc is the observed time of the
occultation center minus the predicted time from the primary
transit, assuming zero eccentricity; r14 = t14(occ)/t14(tra),
that is the ratio of the secondary and primary transit durations.
Assuming that the errors are independent on the transit times
and durations both for the primary and the secondary transits
and that these errors are also uncorrelated with the error of
the period, we can use the above equation to estimate the ec-
centricity and its pure statistical error. For the primary transit
and period we take the values given in Table 4 of Delrez et
al. (2016). For the secondary transit we use the values shown in
Table 3 of this paper. Errors are assumed to be Gaussian. Then,
Eq. 6 yields e = 0.0207 ± 0.0153 if we use the SMARTS and
e = 0.0314±0.0222, if we use the TRAPPIST data. These eccen-
tricity values are also tested by using the primary transit center
values of Evans et al. (2018) for the HST/STIS G430Lv2 band
(we get very similar results also for the other bands). We note
that this test is not entirely consistent, since we use the transit du-
ration value of Delrez et al. (2016), because, Evans et al. (2018)
do not give this parameter for their data. We get for the SMARTS
and TRAPPIST data, respectively, e = 0.0198 ± 0.0157 and
e = 0.0312 ± 0.0224, i.e., very close to those estimated on the
basis of the primary transits of Delrez et al. (2016).
In concluding, we note that Delrez et al. (2016) give a 3σ up-
per limit of e = 0.07 from the global analysis of the photometric
and radial velocity data. Our independent analysis is quite con-
sonant with theirs.
4. Comparison with planet atmosphere models
As of the time of this writing, there are the following secondary
eclipse observations available for WASP-121b. The Sloan z' data
at 0.9 µm by Delrez et al. (2016), the HST data in 1.1 − 1.6 µm
and the Spitzer data at 3.6 µm, both by Evans et al. (2017). The
main panel in Figure 7 shows the two single-band data points
together with our occultation depth in the K band at 2.2 µm (see
also Table 4 for the actual numerical values used). The data are
overplotted on the recent planetary atmosphere models of Evans
et al. (2017) and Parmentier et al. (2018). We note that although
the "No dissociation" model shows very clearly that one has to
consider element dissociation in modeling HJ atmospheres, it is
unphysical, and it is shown merely for exhibiting the extreme
case of neglecting this important physical process. This model
was constructed by using chemical equilibrium chemistry in the
atmospheric structure modul of the global circulation model, but
H2O abundance was fixed in computing the spectrum. However,
the model labelled "Solar composition" is consistent in this re-
spect, and shows that the currently available data are in an over-
all agreement with it6, without making any special assumption
or adjustment. Unfortunately, the situation is somewhat more in-
volved, as there are several other possibilities yielding spectra
rather similar to that of the "Solar composition" model. For ex-
ample, one may increase the heavy metal content of the "Solar
composition" model by a factor of three, without any essential
effect on the emission spectrum -- see Parmentier et al. (2018)
for further details.
Following Winn (2010), by omitting the negligible inclina-
tion effect, we use the following formula to estimate the eccen-
6 By admitting the existence of systematic differences for the HST
near infrared measurements of Evans et al. (2017) -- see inset of Fig 7.
5
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
]
t
p
p
[
/
s
F
p
F
6
5
4
3
2
1
0
No dissociation [Par18]
Solar composition [Par18]
Black body 2700K
Black body 2360K
Retrieved [Eva17]
HST/WFC3
2.0
1.5
1.0
1
2
1.0
2.0
1.5
4
3
Wavelength [µm]
Fig. 7. Comparison of the single-band secondary transit depths
with the planetary atmosphere models of Parmentier et al. (2018,
[Par18]) and Evans et al. (2017, [Eva17]). The error bars show
3σ statistical uncertainties. We warn that the "No dissociation"
model is unphysical, and is shown merely to limelight the effect
of omitting dissociation in the computation of the spectrum (see
text for further details). The black-body lines correspond to dif-
ferent efficiency of the day/night heat transport (black: fully effi-
cient; gray: no heat transport). For completeness, the inset shows
the HST observations of Evans et al. (2017) with their spectrum
retrieval model and the solar composition model of [Par18]. For
better visibility, we use 1σ error bars here.
Table 4. Secondary transit depths of WASP-121b.
Instr./Filter
TRAPPIST (z')
2MASS K
Spitzer/IRAC
λc
0.9
2.2
3.6
Depth
0.697
2.280
3.670
Source
Error
0.081 Delrez et al. (2016)
0.230
0.130
Evans et al. (2017)
this paper
Notes:
Only single-band data are shown. -- The central wavelength λc is given
in µm, the depth and its 1σ statistical error are given in ppt (part per
thousand).
a deviation of 3.2σ (the other two points deviate less than 0.5σ).
Repeating the same comparison for the retrieved model of Evans
et al. (2017), we get, respectively, 0.238 ppt and 4.7 for the RMS
and χ2. Now all points deviates just barely under 1σ, except for
the 2.2 µm point, that deviates by 1.6σ.
It is important to note that the status of the outliers might
change with a different way of handling systematics. As men-
tioned, over-correcting systematics may lead to lower occulta-
tion depth (e.g., we got larger depth from the 0.9 µm data by
∼ 0.1 ppt than the one derived by Delrez et al. 2016, quite likely,
because of the lack of polynomial correction in our derivation).
With the data available today, it seems that the retrieval
model of Evans et al. (2017) is capable to catch most of the
features of the observed spectrum. The fact that our data devi-
ates by 1.6σ from their model spectrum, indicates that although
additional fine tuning is needed, the basic characteristics of the
data are well-matched. On the other hand, the required VO abun-
dance is some thousand times of the solar value, which warrants
some caution (see Evans et al. 2017 and Parmentier et al. 2018
for further discussion of this issue with the emission spectrum).
Additional complications come from the more extensive data
available from HST and ground-based transmission spectrum
measurements. The recent analysis of these data by Evans et
al. 2018) lends further support for a high (10 -- 30-times solar) VO
abundance and lack of T iO. Furthermore, these data also pose
some challenges in explaining the steep rise of the absorption in
the near ultraviolet regime. The authors invoke sulfanyl (S H) as
a possible absorber, since the standard explanation by Rayleigh
scattering fails in the case of WASP-121b, due to the high atmo-
spheric temperature implied by Rayleigh scattering only.
Unfortunately, the currently available data on WASP-121b
still too sparsely populate the more easily measurable part of the
emission spectrum. In the waveband between 2 µm and 4 µm
(where the CO and H2O emissions are the most pronouncing)
additional data would be of great help. High S/N measurements
carried out by instruments like CRIRES at VLT would be clearly
capable to map this crucial region. In addition to the determi-
nation of the abundances of the molecules above, this would
perhaps constrain also the abundances derived from the shorter
wavelength part of the spectrum, where high S/N data gathering
is more difficult.
The black body lines (gray and black) in Figure 7 show
the effect of heat transport from the day-side to the night-
side. Assuming zero Bond albedo in both cases, the black line
displays the case of heat transport with maximum efficiency
(AB = 0, ε = 1.0 -- see Cowan & Agol 2011, Lopez-Morales
& Seager 2007). It is clear that all available data exclude this
possibility and support circulation models that are rather inef-
ficient, resulting in a higher day-side temperature. For WASP-
121b, this temperature seems to be close to 2700 K, correspond-
ing to ε = 0.57, assuming AB = 0. In a comparison with the
models of Evans et al. (2017) -- who use also the above planet
temperature -- we find that their 'retrieved' model slightly under-
estimates our occultation depth by 1.6σ, whereas the mismatch
for the black body line of 2700 K is only 0.4σ.
By scanning the planet temperature, we find that the best-
fitting black body model to the three single-band data points
(weighted equally) is reached when Tp = 2652 K. The RMS
and the χ2 value of the residuals is 0.194 ppt and 15.4, respec-
tively. All points are within 1σ, except for the one at 0.9 µm, that
deviates by 3.8σ. For the solar composition model of Parmentier
et al. (2018) we get 0.244 ppt and 10.3 for the RMS and χ2, re-
spectively. These large values result from the Spitzer data, with
5. Conclusions
We presented the first secondary transit measurements of an ex-
trasolar planet in the near infrared by using a 1-m class tele-
scope. With the ANDICAM imager attached to the 1.3 m tele-
scope of the SMARTS Consortium, we detected an occultation
depth of (0.228 ± 0.023)% in the 2MASS K band from observa-
tions made in three nights on the very hot Jupiter WASP-121b.
We compared this value with theoretical planetary spectra of
Parmentier et al. (2018) and Evans et al. (2017) and found that it
fits perfectly the former model, using solar composition, atmo-
spheric circulation and molecular dissociation. However, when
considering all available secondary transit data (Sloan z', HST
and Spitzer data -- see Delrez et al. 2016 and Evans et al. 2017),
it seems that the VO-enhanced model of Evans et al. (2017) is
preferred over the solar composition model, albeit with a less
favorable match to our data. Although the 2700 K black body
line yields also an acceptable overall fit to the available data, the
more detailed HST spectrum is not reproduced well. Additional
data in the (2 -- 4) µm regime would be very useful to verify
model predictions on CO and H2O emissions and build a more
coherent planet atmosphere model.
6
Kovacs, G. & Kovacs, T.: The secondary transit of WASP-121b
acquisition period is much appreciated. We also thank the referee for the critical
notes on our early interpretation of the planet atmosphere models. The observa-
tions have been supported by the Hungarian Scientific Research Fund (OTKA,
grant K-81373). TK acknowledges the support of Bolyai Research Fellowship.
Additional grants (PD 121223 and K 129249) from the National Research,
Development and Innovation Office are also acknowledged.
References
Adams, A. D. & Laughlin, G. 2018, AJ, 156, 28
Alonso, R. 2018, Handbook of Exoplanets, ISBN 978-3-319-55332-0, id.40
(arXiv:1803.06204)
Anderson, D. R., Temple, L. Y., Nielsen, L. D. et al. 2018, submitted to MNRAS,
(arXiv: 1809.04897)
Bakos, G. ´A., Torres, G., P´al, A. et al. 2010, ApJ, 710, 1724
Baskin, N. J., Knutson H. A., Burrows, A., et al. 2013, ApJ, 773, 124
Cowan, N. B. & Agol, E. 2011, ApJ, 729, 54
Croll, B., Albert, L., Jayawardhana, R. et al. 2015, ApJ, 802, 28
Cruz, P., Barrado, D., Lillo-Box, J. et al. 2015, A&A, 574A, 103
Deeg, H. J. & Tingley, B. 2017, A&A, 599, A93
Delrez, L., Santerne, A., Almenara, J.-M. et al. 2016, MNRAS, 458, 4025
Deming, D., Seager, S., Richardson, L. J., & Harrington, J. 2005, Nature, 434,
740
Evans, T. M., Sing, D. K., Wakeford, H. R. et al. 2016, ApJ, 822, L4
Evans, T. M., Sing, D. K., Kataria, T. et al. 2017, Nature, 548, 58
Evans, T. M., Sing, D. K., Goyal, J. M. et al., 2018, AJ, 156, 283
Kreidberg, L., Line, M. R., Bean, J. L., et al. 2015, ApJ, 814, 66
Komacek, T. D., & Showman, A. P. 2016, ApJ, 821, 16
Lopez-Morales, M. & Seager, S. 2007, ApJ, 667, L191
Martioli, E., Col´on, K. D., Angerhausen, D. 2018, MNRAS, 474, 4264
Parmentier, V., Line, M. R., Bean, J. L. et al. 2018, A&A, 617, 110
Pollacco D. L., Skillen, I., Collier Cameron, A. et al. 2006, PASP, 118, 1407
Stevenson, K. B., Harrington, J., Fortney, J. J., et al. 2012, ApJ, 754, 136
Winn, J. N. 2010, arXiv:1001.2010v5
Zhou, G., Bayliss, D. D. R., Kedziora-Chudczer, L. et al. 2015, MNRAS, 454,
3002
]
t
p
p
[
s
b
o
δ
3.0
2.0
1.0
0.0
0.0
1.0
δcal [ppt]
2.0
Fig. 8. Observed vs calculated secondary transit depths for the
32 extrasolar planets known today with emission measurements
at ∼ 2.2 µm. Nearly all observations lie above the equality line,
corresponding to the calculated/expected black body value, as-
suming effective heat transport from the day side to the night
side. WASP-121b is shown as a red square. Error bars show 1σ
statistical errors.
studies
In agreement with other
(e.g., Adams &
Laughlin 2018, and references therein), our data support
the lack of efficient day-to-night side heat transport (see Fig. 7).
This conclusion is further strengthened if we compare the
predicted and observed occultation depths by using all available
data today. Based on the list of Alonso (2018), we collected
the secondary transit depths measured in the 2MASS K band
for 32 Hot Jupiters (see Croll et al. 2015, Cruz et al. 2015,
Zhou et al. 2015, Martioli et al. 2018 and this paper). The
observed depths as a function of the expected value (assuming
zero Bond albedo and fully efficient heat transport) are shown
in Fig. 8. The figure clearly shows a nearly uniform offset, with
no apparent dependence on the expected depth. The effect is
exacerbated if we consider more realistic albedos as suggested
by recent analyses of full orbit phase curves -- see Adams &
Laughlin (2018).
We arrive to a similar conclusion if we examine the differ-
ence between the observed and calculated occultation depths
as a function, e.g., of the temperature at the substellar point.
Therefore, -- admitting the need for a more complete charac-
terization of the heat distribution by directly measuring the
night- and day-side fluxes (i.e., Komacek & Showman 2016)
-- from the 2.2 µm measurements alone, there does not seem
to exist a correlation between the heat redistribution efficiency
and planet temperature (i.e., Cowan & Agol 2011, Komacek &
Showman 2016). Supporting our result, it is interesting to note
that a similar study by Baskin et al. (2013), based on Spitzer
3.6 µm and 4.5 µm data has led to the same conclusion.
Although our observations were made in a single waveband,
they yield a reasonably solid piece of information both on the
orbital and on the atmospheric characterization of the WASP-
121 system. Together with future emission data in the (2 -- 4) µm
band they will allow to prove or deny the existence of the CO,
H2O emission feature on the day side predicted by the models in
this waveband.
Acknowledgements. We thank to Laetitia Delrez for sending us the secondary
transit observations presented in the discovery paper on WASP-121. We are
grateful to Vivien Parmentier for making the relevant planet atmosphere models
accessible to us and helping in the comprehension of the models. The profes-
sional help given by the SMARTS staff at the Yale University during the data
7
|
1312.2676 | 1 | 1312 | 2013-12-10T05:54:45 | Dispersion in Neptune's Zonal Wind Velocities from NIR Keck AO Observations in July 2009 | [
"astro-ph.EP"
] | We report observations of Neptune made in H-(1.4-1.8 {\mu}m) and K'-(2.0-2.4 {\mu}m) bands on 14 and 16 July 2009 from the 10-m W.M. Keck II Telescope using the near-infrared camera NIRC2 coupled to the Adaptive Optics (AO) system. We track the positions of 54 bright atmospheric features over a few hours to derive their zonal and latitudinal velocities, and perform radiative transfer modeling to measure the cloud-top pressures of 50 features seen simultaneously in both bands.
We observe one South Polar Feature (SPF) on 14 July and three SPFs on 16 July at ~65 deg S. The SPFs observed on both nights are different features, consistent with the high variability of Neptune's storms.
There is significant dispersion in Neptune's zonal wind velocities about the smooth Voyager wind profile fit of Sromovsky et al., Icarus 105, 140 (1993), much greater than the upper limit we expect from vertical wind shear, with the largest dispersion seen at equatorial and southern mid-latitudes. Comparison of feature pressures vs. residuals in zonal velocity from the smooth Voyager wind profile also directly reveals the dominance of mechanisms over vertical wind shear in causing dispersion in the zonal winds.
Vertical wind shear is not the primary cause of the difference in dispersion and deviation in zonal velocities between features tracked in H-band on 14 July and those tracked in K'-band on 16 July. Dispersion in the zonal velocities of features tracked over these short time periods is dominated by one or more mechanisms, other than vertical wind shear, that can cause changes in the dispersion and deviation in the zonal velocities on timescales of hours to days. | astro-ph.EP | astro-ph |
Dispersion in Neptune's Zonal Wind Velocities from NIR
Keck AO Observations in July 2009
Patrick J. Fitzpatrick1, Imke de Pater1,2, Statia
Luszcz-Cook1,3, Michael H. Wong1, Heidi B.
Hammel4
Abstract
We report observations of Neptune made in H-(1.4-
1.8 µm) and K'-(2.0-2.4 µm) bands on 14 and 16 July
2009 from the 10-m W.M. Keck II Telescope using the
near-infrared camera NIRC2 coupled to the Adaptive
Optics (AO) system. We track the positions of 54 bright
atmospheric features over a few hours to derive their
zonal and latitudinal velocities, and perform radiative
transfer modeling to measure the cloud-top pressures
of 50 features seen simultaneously in both bands.
We observe one South Polar Feature (SPF) on 14
July and three SPFs on 16 July at ∼65◦S. The SPFs
observed on both nights are different features, consis-
tent with the high variability of Neptune's storms.
There is significant dispersion in Neptune's zonal
wind velocities about the smooth Voyager wind profile
fit of Sromovsky et al., Icarus 105, 140 (1993), much
greater than the upper limit we expect from vertical
wind shear, with the largest dispersion seen at equa-
torial and southern mid-latitudes. Comparison of fea-
ture pressures vs. residuals in zonal velocity from the
smooth Voyager wind profile also directly reveals the
dominance of mechanisms over vertical wind shear in
causing dispersion in the zonal winds.
Vertical wind shear is not the primary cause of the
difference in dispersion and deviation in zonal veloc-
Patrick J. Fitzpatrick,
Michael H. Wong, Heidi B. Hammel
Imke de Pater, Statia Luszcz-Cook,
Email: [email protected] (PJF)
1Department of Astronomy, University of California, Berkeley,
CA 94720, USA
2Faculty of Aerospace Engineering, Delft University of Technol-
ogy, 2629 HS Delft, and SRON Netherlands Institute for Space
Research, 3584 CA Utrecht, The Netherlands
3Astrophysics Department, American Museum of Natural His-
tory, Central Park West at 79th Street, New York, NY 10024,
USA
4AURA, 1212 New York Ave. NW, Suite 450, Washington, DC
20005, USA
ities between features tracked in H-band on 14 July
and those tracked in K'-band on 16 July. Dispersion in
the zonal velocities of features tracked over these short
time periods is dominated by one or more mechanisms,
other than vertical wind shear, that can cause changes
in the dispersion and deviation in the zonal velocities
on timescales of hours to days.
Keywords infrared: planetary systems; planets and
satellites: Neptune: atmospheres
1 Introduction
The zonal wind velocities of the giant planets can
be derived by tracking the motions of cloud features
in their atmospheres. Accurate tracking of the mo-
tions of cloud features in Neptune's atmosphere was
first achieved with data from the Voyager 2 space-
craft in 1989 (Stone & Miner 1989). Sromovsky et al.
(1993) made a smooth fit to the zonal velocities vs.
latitudes of discrete cloud features in Neptune's at-
mosphere which were tracked by Limaye & Sromovsky
(1991) in Voyager 2 images. Neptune's canonical zonal
wind profile is this smooth Voyager wind profile. The
wind velocities derived from individual features in Nep-
tune's atmosphere since Voyager 2 are observed to re-
main consistent with this smooth Voyager wind pro-
file, with the exception of features which display sig-
nificant deviation in zonal velocity, presumably due
to mechanisms such as vertical wind shear, wave phe-
nomena, or eddy motions, and which have sometimes
been associated with structures such as dark spots (e.g.
Hammel & Lockwood 1997; Sromovsky et al. 2001b,c,
2002). Sromovsky et al. (2001c) found deviations from
the smooth Voyager wind profile in 1998 HST ob-
servations which were consistent with observations in
2
1995 and 1996 (Sromovsky et al. 2001b), but they con-
cluded that more measurements are needed to confirm
a change in Neptune's zonal wind structure.
Striking dispersion and variation in zonal wind veloc-
ities has been observed on Neptune since the Voyager
era (Smith et al. 1989). Limaye & Sromovsky (1991)
found significant dispersion in zonal wind velocities,
with the greatest dispersion found for more short-
lived features, and at equatorial and mid-latitudes.
Sromovsky et al. (1993) also noted significant devia-
tion in the zonal velocities of cloud features about their
smooth Voyager profile fit to the data of Limaye & Sro-
movsky (1991). Hammel & Lockwood (1997) studied
the zonal motions of cloud features in 1995 HST images,
along with those for features found in Voyager images
(Limaye & Sromovsky 1991; Hammel 1989), Voyager
radio occultation data (Lindal et al. 1990), 1991 HST
data (Sromovsky et al. 1995), and ground-based data
(see Sromovsky et al. (1993) Table VII) and noted that
all measurements show dispersion in velocities in nar-
row bands of latitude, which the authors indicated as
evidence for shear, wave phenomena, or other local dis-
turbances. Sromovsky et al. (2001b) found significant
deviation in zonal velocities from the smooth Voyager
wind profile for features tracked in HST data from 1994,
1995, and 1996. The authors found deviation in 1996
HST data to be mostly associated with a Great Dark
Spot at 32◦N, thought to be the same dark spot seen in
1994 HST images of Neptune at 30◦N (Hammel et al.
1995). The authors associated dispersion in this region
with waves propagating from this Great Dark Spot or
associated standing waves.
The more recent analysis of Martin et al. (2012) re-
veals significant dispersion in zonal wind velocities from
the smooth Voyager wind profile. Martin et al. (2012)
observed Neptune in H-band with the Keck AO sys-
tem for 4 hours on UT 20 and 21 August, and for ∼1
hour on UT 1 September, 2001. The authors reliably
measured the relative velocities of almost 200 clouds in
Neptune's atmosphere, characterizing the dispersion in
Neptune's zonal wind velocities about the smooth Voy-
ager wind profile. The authors found significant disper-
sion in zonal wind velocity (with variations as high as
∼500 m/s), greater than the upper limit they placed on
the contribution to dispersion caused by vertical wind
shear, and which they attributed primarily to outlying
transient clouds that do not move with their local mass
flow.
We conduct a similar study as Martin et al. (2012),
except we track the motions of Neptune's cloud fea-
tures in both H- (∼1.6µm) and K'- (∼2.2µm) bands
(in observations separated by ∼3 Neptune rotations),
which are sensitive to a different range of altitudes in
Neptune's atmosphere, to search for potential differ-
ences between the two wavelengths. We also perform
radiative transfer modeling to estimate the pressures
of features. Comparison of the zonal velocities of fea-
tures with their estimated pressures provides a direct
probe of the relative contribution of vertical wind shear
to dispersion in the zonal wind velocities. Because our
observations in H-band probe a greater overall mag-
nitude and range of depths in Neptune's atmosphere
than those in K'-band, comparing the dispersion ob-
served between the two filters can also give us insight
into the relative contribution of vertical wind shear to
dispersion in the zonal winds.
In Section 2 we describe our observations and data,
including our method of alignment and cylindrical pro-
jection of images. In Section 3 we describe our method
of tracking the longitude-latitude positions of atmo-
spheric features in our images. Section 4 describes
our results for the dispersion in Neptune's zonal winds,
along with our results for the depths of features from
radiative transfer modeling, and our observations of the
South Polar Features. In Section 5 we discuss our re-
sults in the context of a few relevant mechanisms which
can cause dispersion in Neptune's zonal winds and we
compare the dispersion observed in H- and K'-bands on
14 and 16 July, respectively. Finally, Section 6 summa-
rizes our conclusions.
2 Data
2.1 Observations and Data Reduction
We observed Neptune from the 10-m W.M. Keck II
Telescope on Mauna Kea, Hawaii, on 14 and 16 July
2009 (UT) as part of a project to study the planet's at-
mosphere and rings at near-infrared (NIR) wavelengths.
H-(1.4-1.8µm) and K'-(2.0-2.4µm) band images were
taken using the narrow camera of the NIRC2 instru-
ment, coupled to the AO system. The 1024×1024 array
has a pixel scale of 9.963±0.011 mas/pixel in this mode
(Pravdo et al. 2006), which at the time of observations
corresponded to a physical scale of ∼210 km/pixel at
disk center.
Although we observed Neptune in both bands on
each day, images were more frequently taken in H-band
on 14 July and in K'-band on 16 July. On 14 July
we took a total of 75 H-band images spanning ∼2.5
hours (11:20:21 - 13:47:44 (UT)) and on 16 July we
took a total of 105 K'-band images spanning ∼3.5 hours
(10:54:57 - 14:20:09 (UT)). The integration times of
both H- and K'-band images are 60 sec, which provide
enough signal to noise without saturating our detector,
3
and assure that smearing due to Neptune's rotation is
less than one image pixel. Short integration times en-
able a dense sampling of images over time, allowing
us to accurately identify the same features in succes-
sive images. Our data were taken in ∼5-min sequences
of five frames each. Largest separations between se-
quences of images were ∼30 min, during which obser-
vations of photometric standards were carried out.
We reduced images using standard infrared data re-
duction techniques of sky subtraction, flat fielding, and
median-value masking to remove bad pixels. All im-
ages were corrected for the geometric distortion of the
array using the 'dewarp' routines provided by P. Brian
Cameron,1 who estimates residual errors at .0.1 pix-
els. We measured angular resolution with the full width
at half maxima (FWHM) of Neptune's moons visible
in our images. We measure FWHM in H-band on 14
July of 0.050±0.004" and in K'-band on 16 July of
0.049±0.005", which are consistent with the diffraction
limit of our telescope at 2µm (van Dam et al. 2004),
and correspond to effective resolutions of ∼1,060 km
and ∼1,037 km at the center of the disk, respectively.
Images were photometrically calibrated using the
star HD201941, and were converted from units of ob-
served flux density to units of I/F, which is defined as
(Hammel et al. 1989a):
I
F
=
r2
Ω
FN
F⊙
,
(1)
where r is Neptune's heliocentric distance, πF⊙ is the
Sun's flux density at Earth's orbit (Colina et al. 1996),
FN is Neptune's observed flux density, and Ω is the solid
angle subtended by a pixel on the detector. By this
definition, I/F=1 for uniformly diffuse scattering from
a Lambert surface when viewed at normal incidence.
2.2 Locations of Cloud Features
Figure 1 shows Neptune images on both days of obser-
vation in H-band (left panels) and K'-band (right pan-
els), taken towards the beginning (1st and 3rd rows)
and end (2nd and 4th rows) of each night. On both
14 and 16 July we observed Neptune at roughly the
same longitudes (these images are separated by ∼3
Neptune rotations). We immediately note that at-
mospheric features tend to be distributed preferen-
tially along bands of constant latitude, with increased
prevalence at mid-latitudes, in agreement with previous
observations (e.g. Sromovsky et al. 2001b; Max et al.
2003; Irwin et al. 2011; Martin et al. 2012). In addition
to clouds at roughly the same latitudes as Martin et al.
(2012), we observe clouds at ∼40◦N. In all
images
there is an absence of cloud features just south of the
equator, in agreement with previous observations (e.g.
Limaye & Sromovsky 1991; Martin et al. 2012). At-
mospheric features vary drastically between 14 and 16
July. Feature morphology changes so dramatically that
we cannot with certainty identify the same features
present on both nights, as has been noted in previous
observations (e.g. Sromovsky et al. 2001b; Karkoschka
2011b).
In all 14 July images we observe a large bright fea-
ture centered at ∼65◦S. In 16 July images, centered
at the same latitude, we observe three distinct fea-
tures instead of the large bright feature seen on 14 July.
Due to foreshortening effects, these three features ap-
pear to begin to coalesce into one as they approach the
limb at the end of the night. Features in this latitude
region have been identified in previous observations
as South Polar Features (SPFs; e.g. Hammel et al.
1989b; Smith et al. 1989; Limaye & Sromovsky 1991;
Sromovsky et al. 1993; Crisp et al. 1994; Hammel &
Lockwood 1997; Sromovsky et al. 2001b,c; Rages et al.
2002; Karkoschka 2011a; Martin et al. 2012). We will
discuss these SPFs in Section 4.4.
In all H-band images we observe a small bright
feature at the south pole of the planet.
In K'-band
we cannot see this feature. This feature has been
observed since the Voyager era (e.g.
Smith et al.
1989; Limaye & Sromovsky 1991; Hammel et al. 2007;
Luszcz-Cook et al. 2010; Karkoschka 2011a). As seen
by Limaye & Sromovsky (1991), this feature persisted
in the Voyager observations over many Neptune rota-
tions, and the authors were tempted to suggest that
it marked the planet's true rotation pole, as doing so
would remove a mean meridional velocity bias which
puzzled them. Martin et al. (2012) assumed this 'south
pole dot' to mark Neptune's true rotation pole in veri-
fying image navigation, and found the planet center de-
duced from the south pole dot to agree with that deter-
mined from limb fitting within 1 image pixel (16.7±0.2
mas). There is no a priori reason to assume the feature
at the south pole to mark Neptune's true rotation pole.
In fact in previous observations a pair of south polar
spots have been seen ∼1-2◦ away from the south pole
(Luszcz-Cook et al. 2010; Karkoschka 2011a), and the
former authors suggested that these clouds may form in
a region of strong convection surrounding a south polar
vortex.
2.3 Image Navigation and Cylindrical Projection
1http://www2.keck.hawaii.edu/inst/post observing/dewarp/
nirc2dewarp.pro
To navigate and align our images we must determine
the location of the physical center of Neptune in each
4
image to within sub-pixel accuracy. We do so using a
multivariate nonlinear χ2 minimization routine which
simultaneously fits for the positions of three moons onto
their respective orbits for each disk image. The orbit
of each moon was derived using the ephemeris genera-
tors in the Rings Node of NASA's Planetary Data Sys-
tem (http://pds-rings.seti.org/). We derive moon or-
bits rather than use their individual locations given by
the ephemeris generators due to uncertainties in the lat-
ter (e.g. Jacobsen & Owen 2004; de Pater et al. 2005).
This is based on the fitting routine used to align Nep-
tune images by Luszcz-Cook et al. (2010). In 14 July
H-band images we fit for the simultaneous positions of
Galatea, Larissa, and Despina, and in 16 July K'-band
images we fit for the simultaneous positions of Galatea,
Larissa, and Proteus. Our mean (x, y) uncertainties in
the derived Neptune centers in our images (calculated
from the variance modified by a factor of the reduced-
χ2) for H- and K'-bands are (0.08 pix,0.07 pix) and
(0.09 pix,0.08 pix), respectively.
The accuracy of the alignment of our images can be
judged from Figure 2, which shows mean averages of
the stack of aligned images in each filter. The positions
of each individual moon trace out visible orbits in the
averaged image. In order to better resolve these orbits
we high-pass filter each averaged image by subtract-
ing from it an identical image that has been median-
smoothed with a width of 30 pixels. High-pass filter-
ing the averaged image eliminates large features that
dominate the intensity range and allows high resolution
of faint structure such as Neptune's moons and rings.
Neptune's Adams and Le Verrier rings are clearly vis-
ible in both images in Figure 2.
In both H- and K'-
bands we can clearly distinguish the orbits of Despina,
Galatea, and Larissa, although those of Despina and
Galatea are difficult to distinguish from the nearby Le
Verrier and Adams rings. In K'-band we can also dis-
tinguish the orbit of Proteus. We superpose the orbits
we derived for these moons in Figure 2. The moon or-
bits physically traced in our averaged high-pass filtered
images align well with the superposed derived moon or-
bits, reflecting the accuracy of our alignment and Nep-
tune center determinations.
After precise location of Neptune's center, we trans-
form the data to regularly gridded, cylindrical coordi-
nates, using the same code described in Asay-Davis et
al. (2009) and Lii et al. (2010). For the transforma-
tion from sky coordinates to planetographic latitude-
longitude coordinates on Neptune's 1-bar surface, we
use equations similar to those in Hueso et al. (2010),
but simplified with the plane-parallel assumption. Fi-
nally, we use IDL's trigrid function to resample the
latitude-longitude data on a regular grid. An example
of a transformed image from our cylindrical projection
is shown in Figure 3.
3 Tracking Atmospheric Features
We use the velocities of cloud features as a tracer for
atmospheric wind velocities, and therefore track the po-
sitions of cloud features in our images. After transfor-
mation of each image frame, we combine the five frames
within each sequence together in a mean average. We
track the positions of cloud features in these trans-
formed averaged images. On average, zonal drift rates
over 5 min are smaller (<0.65◦/5 min) than an effec-
tive angular resolution element at disc center (∼2.4◦).
Averaging images does not significantly smear features
and increases signal-to-noise, allowing us to better dis-
tinguish faint, fine features. Averaging sets of H-band
data from 14 July yields 15 averaged images and av-
eraging sets of K'-band data from 16 July yields 21
averaged images.
In order to track cloud features we first identify them
between successive transformed averaged images by
constructing images such as Figure 4, which shows suc-
cessive transformed averaged images in strips of fixed
latitude (from 52-30◦S) stacked with time increasing
along the vertical axis.
We define a feature as having an observed brightness
distribution distinct from features around it and per-
sisting in at least 4 successive images. Studying Figure
4: bright atmospheric features show a broad range of
dynamics. Some features persist relatively consistently
in brightness and morphology (blue dashed lines), while
others are very ephemeral (yellow dashed lines), ap-
pearing and disappearing or significantly varying in
morphology on minute timescales. It appears that the
smallest features tend also to be the most ephemeral, as
was also noted in the Voyager era (Smith et al. 1989).
Our initial interpretation of images such as Figure 4 is
that features in K'-band evolve more rapidly and ap-
pear more ephemeral. This, however, could be due to
the greater number of smaller features observed on 16
July. We compare the dispersion in zonal velocities
between 14 July H- and 16 July K'-band features in
Section 5.2.
After we identify distinct features, we measure their
longitude-latitude positions in each transformed av-
eraged image following the method of Martin et al.
(2012): for a single feature in a transformed averaged
image we take an initial contour of the image at an in-
tensity level that outlines that feature. After isolating
this feature we then take three more contours of this fea-
ture at intensity levels defined at 60%, 70%, and 80% of
5
the maximum intensity within the initial contour. We
define the center of a feature as the midpoint between
longitudinal and latitudinal extrema of a given contour;
three individual feature center positions are measured
using these three contours. We define the final feature
center position as the average of these three measured
center positions. The uncertainty in feature center po-
sition is defined as the sum in quadrature of the stan-
dard deviation of the three measured center positions
and navigation uncertainty associated with location of
Neptune center position in untransformed images. We
repeat this procedure for each tracked feature through-
out each of the images in which it is identified. An
example illustrating our method is shown in Figure 5.
4 Results
4.1 Deriving Zonal Drift Rates
Our results for 14 July H-band features are shown in
Figure 6 and those for 16 July K'-band features are
shown in Figure 7. Here we show the longitude posi-
tions of each tracked feature versus time, separated into
latitude bins identified above each panel.
At fixed latitude, we expect the longitude positions
of atmospheric features to move linearly with time,
in agreement with the smooth Voyager wind profile
of Sromovsky et al. (1993). We expect the latitudi-
nal speeds of most features to be consistent with zero.
We derive longitudinal and latitudinal drift rates for
each feature by fitting lines to their longitudes vs. time
and latitudes vs. time using a Monte-Carlo iteration
routine. Our routine fits position vs. time data to a
line using the function ladfit in IDL (which fits data to
a linear model using a "robust" least absolute devia-
tion method) with each iteration sampling each posi-
tion measurement from a normal distribution centered
on the original position measurement with a width the
size of the uncertainty in that measurement. We make
∼1,000 iterations until our fits converge. We use the
means of the fit parameters from all iterations as our
output fit parameters, and the standard deviations of
the fit parameters from all iterations as the uncertain-
ties in our output fit parameters.
Our derived drift rates are shown as solid and dotted
lines over our data in Figures 6 and 7. Also shown in
Figures 6 and 7, overplotted onto each feature using a
dashed line, are slopes indicating the longitudinal drift
rates expected at the latitude of each feature according
to the smooth Voyager wind profile of Sromovsky et al.
(1993). We see that, although many features follow
constant drift rates, as we expect, there are also many
features whose individual longitude positions show sig-
nificant variation off of these constant drift rates. Non-
constant velocities could include true variability, but
could also include measurement errors. Sources of er-
ror include changes in cloud morphology, image navi-
gation errors, and feature centroiding errors. For in-
stance, anomalous longitude position measurements at
∼70 min on July 14 at multiple latitudes argue strongly
for some type of measurement error. Therefore, in order
to disentangle the effect of large variations in individ-
ual longitude position measurements, either real or due
to error, from dispersion in derived mean longitudinal
drift rates, we separate features by the mean of their ab-
solute residuals in individual longitude position about
their derived longitudinal drift rates, σrl. Our selection
is shown in Figure 8. We divide features into "Low-
σrl" (σrl ≤ 0.7 deg) and "High-σrl" (σrl > 0.7 deg).
The best-defined feature tracks are in the Low-σrl bin,
which contains 25 of 41 July 14 H-band features, and
29 of 46 July 16 K'-band features.
A combination of true variability and unknown
sources of uncertainty manifests itself as deviations
from linear motion.
In order to obtain upper limits
to uncertainties for the derived zonal velocities of Low-
σrl features, we assume all scatter in the individual
measurements of Low-σrl features about their smooth
motion is due to unknown errors, and do the following:
for each Low-σrl feature whose value for the reduced-
χ2 of observed longitude-time data about its derived
zonal drift rate is greater than 1, we solve for an ad-
ditional contribution to the total uncertainty in longi-
tude position such that the reduced-χ2 is equal to 1,
assuming this additional source of uncertainty to con-
tribute equally to each longitude position measurement
for a given feature, and to be random and uncorrelated
with the sources of uncertainty in longitude position
already considered; that is, for each Low-σrl feature
with χ2
red >1 we solve the following expression for an
additional unknown source of uncertainty in longitude
position, σu:
(φi−φexp,i)2
σ2
cent,i+σ2
nav,i+σ2
= 1,
u
red = 1
χ2
N −2
N
P
i=1
where N is the total number of longitude-time mea-
surements, φi are the measured longitude positions,
φexp,i are the corresponding longitude positions ex-
pected from a feature's derived zonal drift rate, σcent,i
is the standard deviation of the three measured feature
center positions composing the mean longitude position
(from the 60%, 70%, and 80% intensity contours), and
σnav,i is the contribution from navigation uncertainty.
Once we solve for σu this way, we include it as a con-
tribution to the uncertainty in each longitude position
measurement of its corresponding feature and recom-
pute that feature's zonal drift rate and its uncertainty
6
in the same way outlined above. The resulting longi-
tude position errors and derived drift rates are those
shown in Figures 6 and 7. We make this correction for
all Low-σrl features. To quantify the relative magni-
tude of σu, we compute the ratio of σu to the initially
estimated sources of longitude position measurement
uncertainty (ηu,i = σu/(σcent,i + σnav,i)) for each fea-
ture. For 14 July H-band features the mean ηu is 9.4
and for 16 July K'-band features the mean ηu is 4.9. We
make the same correction for errors in latitude position
for all Low-σrl 14 July H-band features and for 25 Low-
σrl 16 July K'-band features. Defining a similar ηu,lat
for unknown errors in latitude position: for 14 July H-
band features the mean ηu,lat is 5.9 and for 16 July
K'-band features the mean ηu,lat is 3.4. All uncertain-
ties in feature velocities and drift rates presented here
are derived including the contribution of σu to measure-
ment uncertainty. Because σu includes both unknown
sources of measurement error and true variability, the
uncertainties we present in feature velocities and drift
rates are upper limits to the true uncertainties.
We distinguish between High- (red) and Low-σrl
(blue) features in Figures 6 and 7. We note that there
is a greater number of High-σrl features found in K'-
band, probably due to the fact that a greater num-
ber of smaller features (which we said tend to be more
ephemeral) were observed on 16 July. We note that a
number of 16 July K'-band features display a pattern
suggestive of east-west temporal oscillation, similar to
what was observed by Martin et al. (2012). However,
this pattern occurs simultaneously at t≃50-150 min and
with similar periods and phases for a number of features
at different locations on the planet. This suggests that
the observed pattern may be caused by some type of
measurement error, as discussed above. For this reason
we also classify these features as High-σrl. We focus on
Low-σrl features in our analysis of dispersion in wind
speeds.
Comparing the drift rates expected from the smooth
Voyager wind profile with our derived drift rates in Fig-
ures 6 and 7 we note deviation from the smooth Voyager
wind profile. Variation in zonal drift rates within indi-
vidual latitude bins reveals dispersion about the smooth
Voyager wind profile. While large deviation from the
smooth Voyager wind profile is more frequently found
among High-σrl features, for which deviation in drift
rate is strongly entangled with scatter in individual po-
sition measurement, Low-σrl features show significant
dispersion as well. We note, however, that we cannot
fully confirm or quantify dispersion in zonal velocities
from Figures 6 and 7, because uncertainties in the de-
rived drift rates are not shown. We better quantify the
dispersion we observe in Neptune's zonal wind veloci-
ties in what follows.
4.2 Dispersion in Zonal Wind Velocities
We translate the drift rates of atmospheric features into
wind velocities according to the following relations:
Vlon = Req cos θ
dφ
dt
,
Vlat = (cid:0)Req sin2 θ + Rpol cos2 θ(cid:1)
dθ
dt
,
(2)
(3)
where Vlon and Vlat are the zonal and latitudinal veloc-
ities of atmospheric features (m/s), Req=24,766×103
m is Neptune's equatorial radius, Rpol=24,342×103
m is Neptune's polar radius, and dφ/dt and dθ/dt
are derived zonal and latitudinal drift rates (rad/s),
where, for dφ/dt, motion from astronomical east to
west is taken to be the positive direction. Our de-
rived zonal velocities for 14 July H- and 16 July K'-
band Low-σrl features are shown in Figures 9 and 10,
respectively, against the smooth Voyager wind profile
of Sromovsky et al. (1993) (black solid line). They are
also listed for each feature in Tables 2 and 3, along with
other relevant quantities associated with the motion of
each feature. We plot features such that the size of the
square used to represent distinct features increases lin-
early with the length of time over which a feature was
tracked. The longest and shortest tracking times we
obtain in H-band are 139 min and 24 min, respectively,
and the longest and shortest tracking times we obtain
in K'-band are 198 min and 21 min, respectively.
Considering both Figures 9 and 10 we observe signif-
icant deviation from and dispersion about the smooth
Voyager wind profile. Low-σrl features which scat-
ter most from the smooth Voyager wind profile are
those which were tracked for shorter time periods. In-
creased dispersion is seen at equatorial and southern
mid-latitudes. On average, greater dispersion is seen
among 14 July H-band features.
Although Low-σrl 16 July K'-band features are on
average found to be consistent with the smooth Voy-
ager wind profile within uncertainties, we find signifi-
cant deviation, ∆Vlon,σ, for 16 July K'-band features
from the smooth Voyager wind profile (including its
width of uncertainty) as high as 290±77 m/s (at the
equator).2 Low-σrl H-band features show a significant
2Two descriptions of the deviation in zonal velocities from the
smooth Voyager wind profile are used throughout. ∆Vlon,σ,
which is presented here, takes into account uncertainty in the
smooth Voyager wind profile. For a feature that is faster than
the smooth Voyager wind profile, ∆Vlon,σ ≡ Vlon−(Vvoy + σvoy),
where Vvoy and σvoy are the zonal velocity and its uncertainty
predicted by the smooth Voyager profile at the latitude of the
7
average absolute deviation of 177±55 m/s. Deviation
from the smooth Voyager wind profile of 14 July H-band
features is found as high as ∼500 m/s, although uncer-
tainties in these measurements are large -- as high as
25% at 23◦S and 50% at the equator. We more closely
examine the difference in dispersion between 14 July H-
and 16 July K'-band features in Section 5.2.
Previous studies tracking the motions of Neptune's
cloud features have found significant dispersion and de-
viation in the zonal velocities. Limaye & Sromovsky
(1991) found variation as high as ∼750 m/s in the zonal
velocities of individual features at fixed latitude, with
the greatest variation at equatorial and mid-latitudes.
When only considering features whose uncertainties in
zonal velocity were <25 m/s, the authors found the
standard deviation of the zonal velocities about the
mean, averaged within 1-degree latitude bins, to be as
high as ∼275 m/s. Sromovsky et al. (2001b) tracked
the motions of bright cloud features in HST data from
1994, 1995, and 1996, and found deviation in zonal
velocities from the smooth Voyager profile as high as
∼175 m/s for features tracked over time periods rang-
ing from ∼1-18 hrs. Martin et al. (2012) found devia-
tion in zonal velocities from the smooth Voyager profile
as high as ∼500 m/s for features tracked over ∼1-4 hrs
in NIR Keck AO images of Neptune in 2001. Using the
velocities of cloud features as tracers for atmospheric
wind velocities, we agree with previous results and ob-
serve significant dispersion in Neptune's wind velocities
about its mean zonal wind profile.
4.3 Cloud Feature Pressures From Radiative Transfer
Modeling
We use radiative transfer modeling to estimate the
pressures of a selection of 50 features that were visi-
ble in both H- and K'-bands from the beginning and
end of both nights. Model spectra are produced using
a 300-layer two-stream radiative transfer code, which
is described in detail in Appendix B of Luszcz-Cook
(2012). We adopt the temperature profile derived by
Fletcher et al. (2010) throughout the atmosphere. We
feature considered. For a feature that is slower than the smooth
Voyager profile, ∆Vlon,σ ≡ Vvoy − σvoy − Vlon. According to
this definition, a positive value of ∆Vlon,σ greater than its un-
certainty represents a feature with significant deviation in zonal
velocity, while a negative value of ∆Vlon,σ represents a feature
that is consistent with the smooth Voyager profile. The quan-
tity ∆Vlon ≡ Vlon − Vvoy measures deviation from the smooth
Voyager wind profile without consideration of uncertainty in the
latter. This quantity is used mainly when comparing deviation
of two features or sets of features. These quantities are specified
in context. A Graphical illustration of these two definitions of
deviation is shown in the top legend of Figure 9.
assume a mixing ratio of 0.15 for He and 0.003 for
N2 (Conrath et al. 1993). The methane (CH4) abun-
dance follows Fletcher et al. (2010) in the upper atmo-
sphere and remains at a mole fraction of 0.022 in the
troposphere below the condensation level (Baines et al.
1995). The gas opacity at these wavelengths is dom-
inated by H2 collision-induced absorption (CIA) and
CH4 opacity; for CIA we use the coefficients for hydro-
gen, helium and methane from Borysow et al. (1985,
1988) and Borysow (1991, 1992, 1993), assuming an
equilibrium ortho/para ratio for H2. For methane, we
use the correlated-k method and follow the recommen-
dations of Sromovsky et al. (2012) for outer planet NIR
spectra.
As the spectral information in our data is limited, we
use a simplified model of the atmospheric cloud/haze
distribution. We assume the presence of an opti-
cally thick bottom 'surface' cloud and set the depth of
this cloud to 2.4 bar. We set the Henyey-Greenstein
asymmetry parameter of the bottom cloud to -0.1
(preferentially backscattering) and adjust the single
scattering albedo to match the data as described in
Luszcz-Cook et al. (2010). Luszcz-Cook (2012) shows
that this model fits the µ dependence of the spectrum in
a dark part of Neptune's disk in field-integral spectro-
scopic (OSIRIS) data obtained at Keck. For this sim-
plified model, we allow for one additional aerosol layer
above the bottom cloud; the particles in this higher-
altitude haze/cloud layer are treated as Mie scatterers:
we assume that ensembles of particles are distributed
according to
n (r) ∝ r6 exp (−6r/rmax)
(4)
where n (r) is the number density of particles of radius
r, and rmax is the maximum in the particle distribu-
tion (Hansen & Pollack 1970) and is set to 1.0 µm (e.g.
Irwin et al. 2011). This particle size distribution is
analogous to that found on Titan (Mitchell et al. 2011).
The extinction cross section and Henyey-Greenstein
asymmetry parameter of the scattering are calculated
using Mie theory. We assume that the particles have a
scale height that is 0.1 times the gas scale height, cor-
responding to physically thin cloud layers (Irwin et al.
2011). The free parameters in the model are the num-
ber density of cloud particles at the bottom (maximum)
pressure of the haze/cloud, and the altitude (pressure)
of the cloud. For each feature identified in the data for
modeling, we fit the observed H- and K'- band I/F val-
ues given the viewing geometry (µ) in the following way:
for each of 40 model pressure levels distributed (loga-
rithmically) from 3 mbar to 3 bar, we determine the
cloud particle number density that would best match
the model H-band I/F to the observed H-band I/F.
8
We repeat this procedure for K'-band, then we find the
pressure at which the H and K'-band best-fit number
densities agree; that is, we determine the pressure at
which a cloud of some particle density can best repro-
duce both the H- and K'-band I/F values.
Cloud-top pressure retrievals are shown in Figures
11a and 11b. At different latitudes, the range of de-
rived cloud pressures varies by as much as a factor of
∼4, with the greatest range in pressures found at south-
ern mid-latitudes (60◦S to 27◦S), spanning ∼0.5 bar
(∼2 scale heights). Clouds at northern latitudes (above
20◦N) lie at higher altitudes (∼0.1-0.2 bar). We note,
however, that clouds at northern latitudes are system-
atically higher in emission angle, and from these lim-
ited data we cannot rule out a systematic bias in the
model with µ. Although the 2D relation of pressure
vs. µ in Figure 11b seems suggestive of a bias with µ,
this effect is entangled with any dependence of cloud
pressure on latitude, and within fixed latitude bins (in-
cluding the 3rd dimension shown by symbol shape and
color in Figure 11b) there is not an obvious bias with µ.
For example, aside from one feature at ∼50◦S found at
∼0.6 bar, clouds at equatorial latitudes (27◦S to 5◦N)
are uniformly found at deeper altitudes (∼0.5 bar) in-
dependent of viewing angle.
Our observation that the northern features appear to
be at the highest altitudes is consistent with previous
authors (e.g. Sromovsky et al. 2001b; Gibbard et al.
2003; Luszcz-Cook 2012). The precise values of the
derived northern cloud pressures do not, however,
agree with these works, which find them in the strato-
sphere at 0.023-0.064 bar. We expect these differ-
ences to be related to the limitations of our data
(we only have broadband measurements, not spec-
tra), differences in the sensitivity of our measure-
ments to different altitudes (for example, Gibbard et al.
(2003) measures mostly in K'-band, which is not
sensitive to altitudes as deep as those we probe in
H-band), and the simplicity of our model -- previ-
ous studies have favored models with a more com-
plicated haze structure and which vary other model
parameters (e.g. Baines & Smith 1990; Gibbard et al.
2002; Irwin et al. 2011; Karkoschka & Tomasko 2011;
Luszcz-Cook 2012). Our finding that equatorial fea-
tures are deepest, while the SPFs in the south are found
above them (∼0.3 bar), is different from the results of
Gibbard et al. (2003), which suggest a trend of increas-
ing altitude with latitude from south to north. How-
ever, clouds at these equatorial latitudes were not ob-
served at the earlier epoch (see Gibbard et al. (2003)).
4.4 South Polar Features
We observe one large bright SPF at 64◦S on 14 July
and three smaller SPFs centered at ∼65◦S on 16 July.
SPFs near these latitudes have been seen since the
Voyager era (e.g.
Smith et al. 1989; Hammel et al.
1989b; Limaye & Sromovsky 1991; Sromovsky et al.
1993, 2001c; Rages et al. 2002; Karkoschka 2011a;
Martin et al. 2012) although they have displayed a
latitudinal shift, mostly occurring at 67-74◦S un-
til 2004, and at 60-67◦S since, consistent with our
observations (Karkoschka 2011a).
Individual SPFs
are rapidly-evolving features with well-defined peri-
ods that cannot be tracked from one planet rotation
to the next, and move through the larger structure
of the SPF formation region (e.g. Smith et al. 1989;
Limaye & Sromovsky 1991; Sromovsky et al. 1993);
these features have been observed to form in a fixed
longitude region rotating at nearly the planet inte-
rior period (Hammel et al. 1989b) -- that inferred from
Voyager's radio data tracking Neptune's magnetic field
(Warwick et al. 1989; Lecacheux et al. 1993), then to
move East with well-defined periods ranging from 11.7
hr at 74◦S to 13 hr at 68◦S, and dissipate before mov-
ing halfway around the planet (Limaye & Sromovsky
1991; Sromovsky et al. 1993). At some times, the SPF
region can be completely free of bright cloud features
(e.g. Sromovsky et al. 1993; Rages et al. 2002). At
some times individual SPFs can sporadically form a
cloud clump that can be the brightest feature on the
disk, with the brightening lasting only tens of hours or
less (Rages et al. 2002; Sromovsky et al. 2001c), as we
observe on 14 July.
At the beginning of the night on 14 July we observe
the large SPF centered at a longitude-latitude position
(152◦W, 64◦S) to move with a zonal velocity of 272±15
m/s, consistent with the smooth Voyager wind profile,
and a latitudinal speed consistent with zero. At the be-
ginning of the night on 16 July we observe three SPFs
at centroid positions (130◦W, 67◦S), (119◦W, 63◦S),
and (92◦W, 65◦S) all moving with zonal velocities con-
sistent with the smooth Voyager wind profile except for
the SPF at 63◦S, which moves slower than the smooth
Voyager profile by ∆Vlon,σ=70±45 m/s. Two 16 July
SPFs are found with significant north-south velocities:
the SPF at 67◦S has a northward velocity of 76±43
m/s, and the SPF at 63◦S has a southward velocity
of 117±53 m/s. If we extrapolate the zonal drift rate
of the 14 July SPF we expect its centroid location at
10:55 on 16 July (UT) to be at a longitude of 40±14◦W,
whereas the longitudinal extent of the 16 July SPFs
does not reach below ∼80◦W at that time. This
strongly suggests that the SPFs observed on the two
dates are different features, and that storms develop
and decay on timescales of hours to days, consistent
with previous observations (e.g. Limaye & Sromovsky
1991; Sromovsky et al. 1993, 2001c; Rages et al. 2002).
9
The dynamics underlying the SPFs have not yet been
fully addressed. Sromovsky et al. (1993) found evi-
dence for strong convection driving the SPFs. Karkoschka
(2011a) found a rotational lock between the SPF for-
mation region and the South Polar Wave (SPW), a
southern n=1 wave spanning the latitudes 65-55◦S, vis-
ible as a dark band in Voyager and HST data (e.g.
Smith et al. 1989; Sromovsky et al. 2001a,b). The au-
thors suggested that the vertical motions causing the
formation of SPFs are dynamically linked to the SPW,
and that the SPW itself is vertically phase-locked with
the planet interior. The authors indeed used the mo-
tions of the SPF formation region and SPW to infer
Neptune's rotational period (15.9663±0.0002 hr), dif-
ferent from the 16.11 hr period measured using Voy-
ager's radio data tracking Neptune's magnetic field
(Warwick et al. 1989; Lecacheux et al. 1993). There
is still much to be understood about the dynamics of
the SPFs, including their temporal variability.
5 Discussion
Here we briefly discuss a few mechanisms that can cause
dispersion in Neptune's zonal winds which can be ad-
dressed by our observations. We then compare the dis-
persion observed in H-band on 14 July with that ob-
served in K'-band on 16 July.
5.1 Sources of Dispersion
We can constrain the contribution of vertical wind shear
to the zonal wind dispersion we observe using our re-
sults for the cloud-top pressures of features. We find a
range in pressures from ∼0.6 bar (at 50◦S) to ∼0.1 bar
(northern features), extending approximately 2 scale
heights. Voyager IRIS observations of temperature as a
function of latitude suggest that at these pressures (30-
1000 mbar) Neptune's vertical wind shear can be on
the order of 30 m/s per scale height, with a maximum
near the equator (Conrath et al. 1989). For a range in
altitudes of 2 scale heights, and assuming that the alti-
tudes of observed cloud features are similar to those for
which we derive the cloud-top pressures, vertical wind
shear should not contribute more than ∼60 m/s to zonal
wind dispersion. If we extend this range of depths up
to the 2.4 bar bottom cloud we assume in our models
and down to 0.02 bar (lower limit for northern features
found by Gibbard et al. (2003)), then feature altitudes
span no more than ∼5 scale heights. For this range of
depths vertical wind shear should not contribute more
than ∼150 m/s to zonal wind dispersion. In many cases
we see greater dispersion in zonal wind velocities than
what is expected from vertical wind shear, even when
considering only Low-σrl features, and especially in H-
band. Vertical wind shear cannot be the only cause of
the dispersion we observe in the zonal velocities.
We can more directly limit the relative contribution
of vertical wind shear to dispersion we observe in the
zonal velocities: Figure 11c shows the cloud-top pres-
sures vs. residuals in zonal velocity from the smooth
Voyager wind profile, ∆Vlon, of Low-σrl features which
were visible in both H- and K'-bands towards the be-
ginning of each night, separated into thin latitude bins
extending from 32◦S to 28◦N -- where, at these alti-
tudes, vertical wind shear is expected to be most im-
portant (see Conrath et al. (1989)). Although we do
not observe features in any single latitude bin over a
wide enough range of pressures to see the vertical wind
shear clearly manifest itself given our uncertainties, Fig-
ure 11c directly indicates that one or more other mech-
anisms dominate over vertical wind shear in producing
dispersion in the zonal velocities: there is significant
zonal dispersion within latitude bins where features are
all estimated to be at about the same pressure (1◦S
to 4◦N and 26-28◦N). The zonal dispersion at fixed
pressure and at fixed latitudes between 1◦S and 4◦N
is greater than even the overall upper limit to zonal
wind dispersion we expect from vertical wind shear.
Features found at 32-29◦S display the only significant
range of pressures (∼0.25-0.45 bar). In this latitude bin
we see a spread in zonal velocities significantly greater
than what is expected over the range from vertical wind
shear (<20 m/s), and we do not observe a trend to-
wards higher speed with depth (which at these lati-
tudes is toward more negative velocities), as would be
expected from vertical wind shear (see Conrath et al.
1989, 1991).
Along with significant variation in zonal wind ve-
locities at fixed latitude, we also observe some signif-
icant north-south feature velocities. Figure 12 shows
the derived north-south velocities of Low-σrl features
against their latitude positions for both 14 July H-
(panel a) and 16 July K'-band (panel b) features. For
a few Low-σrl features we find north-south velocities
that are convincingly significant, and as high as ∼200
m/s. We find that a few features which display sig-
nificant north-south velocities also display significant
deviation in zonal velocities from the smooth Voy-
ager wind profile. Noting significant examples:
the
14 July H-band feature at the equator with residual
zonal speed Vlon,σ=480±254 m/s has a southward wind
speed 215±85 m/s. A 14 July H-band feature at 7◦N
is found with residual zonal speed 180±24 m/s and
northward speed 148±15 m/s. We find a 16 July K'-
band feature at 37◦S with a northward wind speed of
10
180±45 m/s and a residual zonal speed of 117±80 m/s.
Again we note the 16 July SPF at 63◦S with south-
ward velocity 117±53 m/s and zonal speed residual
70±45 m/s. Significant north-south velocities and de-
viations in zonal velocity from the mean zonal profile
indicate the presence of one or more mechanisms which
can cause both north-south and east-west residual mo-
tions.
In particular, although we do not confirm the
presence of temporal oscillation in our observations, we
cannot rule out vortices causing the dispersion we ob-
serve. Cloud features centered on or with motions near
a vortex would display significant north-south and east-
west temporal oscillations, and if these had sufficiently
long periods the motions would appear linear on these
timescales, with a slope different from the smooth Voy-
ager profile, as we observe. Vortices have often been
associated with the dark spots observed on Neptune in
Voyager images (such as Voyager GDS and DS2; e.g.
Polvani et al. 1990; Sromovsky et al. 1993) and in HST
images (such as NGDS-32; e.g. Hammel et al. 1995;
Sromovsky et al. 2001b, 2002).
The observed dispersion in zonal velocities and sig-
nificant north-south velocities imply that we cannot
rule out the contribution of wave mechanisms with suf-
ficiently long periods, such as Rossby waves. We do not
observe the presence of wave mechanisms which oscil-
late with periods on the order of our observing period.
In particular, we do not observe evidence of tidal forc-
ing by Triton, as was suggested by Martin et al. (2012).
Even if we were to assume the motions of those features
we noted seemed suggestive of east-west temporal os-
cillation to be real and not due to error, their periods
would not be near the M2 period of tidal forcing by
Triton of 7.24 hrs.
From ground-based spectroscopic observations prior
to the Voyager encounter, Baines & Smith (1990) found
evidence that dynamically driven sublimation and con-
densation resulting from vertical motions in the at-
mosphere are responsible for rapid changes observed
in Neptune's clouds. Limaye & Sromovsky (1991) also
considered this in the context of the dispersion they
observed in the zonal velocities of features in Voyager
images. The analysis of Martin et al. (2012) supported
this conclusion. The authors noted that if feature ve-
locities represent true fluid velocities, for the fluid to
remain nearly divergence-free (for sub-sonic flow), the
variation they observed in the zonal velocities with east-
west distance (>400 m/s over ∼20,000 km) would im-
ply large east-west gradients in north-south velocities,
which they did not observe, or large vertical motions
incompatible with the atmosphere's gradient Richard-
son number at the observed altitudes. They therefore
confirmed that at least some of the dispersion in zonal
velocities they observed is due to transient clouds which
do not move with the flow. Our observations also sup-
port this result. The spatial distribution of variation
in velocities we observe implies that at least some of
the dispersion is due to transient clouds, and evidence
of cloud motions due to dynamically driven sublima-
tion and condensation is provided by our observations
of clouds that are very ephemeral and rapidly change
morphology.
5.2 Comparing Dispersion
The results shown in Figures 9 and 10 are immediately
suggestive of an overall difference in the magnitude of
dispersion in zonal wind velocities about the smooth
Voyager wind profile between features tracked in H- and
K'-bands on 14 and 16 July, respectively. Whereas 16
July K'-band features seem to agree reasonably well
with the smooth Voyager wind profile, 14 July H-band
features appear to show overall greater dispersion and
deviation from the smooth Voyager profile. The differ-
ence in zonal dispersion and deviation between H- and
K'-bands might be expected to be a result of the greater
range in altitudes probed in H-band over K'-band. Fea-
tures in H-band can be seen down to larger depths than
those in K'-band because the strong absorption of H2
and CH4 in K'-band limits detection of features only
to higher altitudes (see Luszcz-Cook et al. (2010) Fig.
4).
If the difference in zonal velocities is due to the
difference in the range of depths probed in H- and K'-
bands, then the likely responsible mechanism is vertical
wind shear. The effects of vertical wind shear in this
difference might be seen in two ways: first, we would
expect to see a greater spread in the zonal velocities
of features tracked in H-band at fixed latitude; second,
and perhaps more subtly, we might expect to see aver-
age shifts in the zonal velocities of features between the
two bands, consistent with features tracked in H-band
on average being at greater depths than those tracked
in K'-band. To test the latter of these possibilities we
fit Low-σrl Vlon-φ data to a polynomial function of the
form Vlon = a + bφ2 + cφ4 to obtain the best-fit values
of a, b, and c using a Levenberg-Marquardt method, for
comparison with the smooth Voyager wind profile fit of
Sromovsky et al. (1993).3 The results of our polyno-
mial fits and the 1σ uncertainties (calculated from the
variance modified by a factor of the reduced-χ2) are
shown with solid and dashed red lines in Figures 9 and
10 and are listed in Table 1, along with the smooth Voy-
ager profile fit of Sromovsky et al. (1993). The zonal
3We note that Voyager and HST observations were mainly at
visible wavelengths and are probably sensing more deeply than
K'-band images.
11
velocities of Low-σrl features measured in K'-band on
16 July agree well with the smooth Voyager wind pro-
file (within 1σ), whereas the zonal velocities of Low-
σrl features measured in H-band on 14 July are best
represented by a profile which is shifted towards more
positive velocities by 180±50 m/s.
Interestingly, this shift is in the opposite direction
from what we would expect due to vertical wind shear.
If we assume that most features shown in Figures 9
and 10 are at similar altitudes as those for which we
measure the cloud-top pressures, or at least that the
zonal direction of the vertical wind shear with pres-
sure is the same at the altitudes of these features as
those to which the results of Conrath et al. (1989) and
Conrath et al. (1991) apply, then as a result of vertical
wind shear we would expect a general decay in zonal
speed with height. We observe a shift of our 14 July H-
band profile fit towards lower zonal speeds (more pos-
itive velocities) at latitudes from ∼30◦S-30◦N, where
the results of Conrath et al. (1989) and Conrath et al.
(1991) show the vertical wind shear is most important,
and where features that most strongly drive the offset
in our profile fit are found. The shift in our profile fit
is in the direction we would expect if 14 July H-band
features were on average found at higher altitudes than
K'-band features. The effects of vertical wind shear are
not found in differences in the average zonal velocities
of 14 July H- and 16 July K'-band features. We note for
clarity that we do not interpret differences in our pro-
file fits to the latitudinal distribution of zonal velocities
of Low-σrl features as evidence for a consistent change
in Neptune's mean zonal wind profile from the smooth
Voyager wind profile -- the short observing times and
latitude gaps in our data make it ill-suited to yield ev-
idence for such a claim. In what follows we continue to
interpret the smooth Voyager wind profile as describing
Neptune's mean zonal wind profile.
If vertical wind shear were the cause of the apparent
difference in zonal dispersion and deviation between 14
July H- and 16 July K'-band features then we would
also expect: 1) greater spread in 14 July H-band fea-
ture zonal velocities over that of 16 July K'-band fea-
ture zonal velocities at fixed latitude; 2) greater devi-
ation from the smooth Voyager wind profile of 14 July
H-band feature zonal velocities over that of 16 July K'-
band feature zonal velocities at fixed latitude (and this
difference in deviation would be fully accounted for by
the difference in spread in zonal velocities); 3) differ-
ences in the deviation and spread in zonal velocities
would not exceed the upper limits we expect for disper-
sion in the zonal velocities due to vertical wind shear
(Section 5.1). To explore this we do the following: we
separate ∆Vlon for 14 July H- (squares) and 16 July K'-
band (triangles) Low-σrl features (shown in Figure 13a)
into latitude bins (shaded bars) chosen according to the
latitude bands along which features are observed to be
centered (see Figure 1). In each of these latitude bins
we compute a weighted average of the absolute value
of residuals in zonal velocity of Low-σrl features from
the smooth Voyager wind profile, mean ∆Vlon, com-
paring features tracked in H- and K'-bands on 14 and
16 July, respectively (we do not consider uncertainty in
the smooth Voyager profile in this calculation, and so
these results are useful mainly as a comparison between
14 July H- and 16 July K'-band features). The results
are shown in Figure 13b. On average, 14 July H-band
features show greater deviation from the smooth Voy-
ager wind profile than K'-band features, and especially
at equatorial and southern mid-latitudes.
To directly examine differences in the dispersion in
zonal velocity, within each latitude bin where two or
more Low-σrl features are found, we compute the dif-
ference between the fastest and slowest Low-σrl zonal
velocities, max δVlon ≡ (max Vlon − min Vlon), compar-
ing between 14 July H- and 16 July K'-band features.
This is shown in Figure 13c.
In many latitude bins
uncertainties in the zonal velocities are too large to de-
termine whether or not differences in mean ∆Vlon be-
tween 14 July H- and 16 July K'-band features can be
fully accounted for by differences in their max δVlon.
Only in the bin of features centered at 30◦S is there
a significant difference in max δVlon that can fully ac-
count for a corresponding significant difference in mean
∆Vlon. However, our comparison in Figure 13 is still
useful for making statements about the relative con-
tribution of vertical wind shear to differences in the
dispersion in zonal velocities: there are differences in
max δVlon between Low-σrl 14 July H- and 16 July
K'-band features in latitude bins centered at 45◦S and
23◦S that are much greater than overall upper lim-
its to dispersion we expect from vertical wind shear.
These indicate the dominance of one or more mecha-
nisms other than vertical wind shear causing differences
in the dispersion in zonal velocities between 14 July H-
and 16 July K'-band features at these latitudes. Max-
imum possible difference in max δVlon in the latitude
bin centered at 37◦S cannot account for the difference
in mean ∆Vlon in the same latitude bin; the maxi-
mum possible increase in max δVlon of these 14 July
H-band features over that of 16 July K'-band features
(calculated at opposite extremes of their uncertainties)
is 172 m/s, whereas 14 July H-band features at these
latitudes display mean ∆Vlon greater than that ob-
served for 16 July K'-band features by >250 m/s. If
we assume that mechanisms which cause deviation in
zonal velocities from the smooth Voyager wind profile
are similarly capable of causing dispersion in zonal ve-
locities, this also indicates the presence of a mechanism
12
Data
Voyager
14 July H-Band
16 July K'-Band
Constant Term (m/s)
-398±12
-205±49
-387±37
φ2 Term (m/s/deg2)
0.188±0.014
0.083±0.06
0.165±0.041
φ4 Term (m/s/deg4)
-1.2E-5±0.3E-5
7.0E-6±14.0E-6
-1.3E-5±1.1E-5
Table 1 Polynomial fits to feature zonal velocities vs. latitudes and their 1σ uncertainties.
other than vertical wind shear which causes differences
in the dispersion in zonal velocities of 14 July H- and
16 July K'-band features.
Our observations indicate the dominance of one or
more mechanisms over vertical wind shear in produc-
ing the overall dispersion we observe in the zonal winds.
That the difference in deviation and dispersion in zonal
velocities observed between 14 July H- and 16 July K'-
band features is, at least at most latitudes between
∼50-20◦S, not primarily attributable to vertical wind
shear further suggests that the mechanisms which dom-
inate dispersion in the zonal winds can cause changes in
the magnitude of dispersion and deviation in the zonal
winds on timescales of hours to days. This is consistent
with the mechanisms discussed above.
6 Summary and Conclusions
In this study we tracked the longitude-latitude positions
of Neptune's atmospheric features seen in Keck AO im-
ages in H-band on 14 July 2009 and in K'-band on 16
July 2009 over time. We derived zonal and latitudinal
drift rates and velocities for these features. We also
performed radiative transfer modeling for features on
both nights which were simultaneously visible in both
H- and K'-bands. The results we find are the following:
1. We find significant dispersion in the zonal velocities
of Neptune's cloud features about the smooth Voy-
ager wind profile of Sromovsky et al. (1993), with
the largest dispersion seen at equatorial and south-
ern mid-latitudes. The observed dispersion is much
greater than the upper limit we expect due to verti-
cal wind shear. Considering only Low-σrl features:
deviation in zonal velocity from the smooth Voy-
ager wind profile is found as high as 290±77 m/s
(at the equator) among features tracked in K'-band
on 16 July, and as high as ∼500 m/s among features
tracked in H-band on 14 July (although uncertainties
for these measurements are as high as 25% at 23◦S
and 50% at the equator). Comparison of the zonal
velocities and cloud-top pressures of features at fixed
latitude also directly indicates the dominance of one
or more mechanisms over vertical wind shear in pro-
ducing dispersion in the zonal winds. Some features
display significant north-south velocities (as high as
∼200 m/s), and a few of these also display significant
deviation in zonal velocity from the smooth Voy-
ager wind profile. This suggests that zonal disper-
sion is partially due to a mechanism which can pro-
duce residual motion along both directions, such as
vortices and wave mechanisms with sufficiently long
periods so as to appear linear on these timescales,
such as Rossby waves. Our observations are consis-
tent with the result that zonal dispersion is caused at
least in part by transient clouds due to dynamically
driven sublimation and condensation.
2. Radiative transfer modeling indicates that, aside
from one feature at 50◦S found at ∼0.6 bar, cloud
features at equatorial latitudes (∼5◦S-5◦N and 27◦S-
15◦S) uniformly lie deeper in the atmosphere (∼0.5
bar) while clouds in the north (above 20◦N) are
found at higher altitudes (∼0.1-0.2 bar). Due to
limitations of our data, differences in the sensitivity
of our measurements to different altitudes, and the
simplicity of our model, the precise altitudes we mea-
sure for the northern features are different from the
results of previous studies, which place them higher,
in the stratosphere.
3. We observe one large SPF at 64◦S on 14 July and
three smaller SPFs centered at about the same lat-
itude on 16 July. Two 16 July SPFs display sig-
nificant north-south velocities, and that with the
largest north-south speed (117±53 m/s southward)
also shows significant deviation from the smooth
Voyager wind profile (70±45 m/s slower). Extrap-
olation of the zonal drift rate of the 14 July SPF
and comparison with the average position and spa-
tial extent of the 16 July SPFs shows that the SPFs
observed on the two nights are different features,
indicating that storms can develop and decay on
timescales of hours to days, in agreement with pre-
vious observations.
4. There is greater dispersion and deviation observed in
the zonal velocities of features tracked in H-band on
14 July than in those tracked in K'-band on 16 July.
Polynomial fits to the zonal velocities vs.
latitudes
of our data show that while 16 July K'-band feature
zonal velocities agree well with the smooth Voyager
wind profile, 14 July H-band feature zonal velocities
are best described by a profile that is shifted to-
ward more positive velocities by 180±50 m/s. This
shift is in the opposite direction from what we would
13
expect if differences in the deviation in zonal veloc-
ities from the smooth Voyager wind profile between
14 July H-band and 16 July K'-band features were
due to vertical wind shear, as a result of greater av-
erage depth of features tracked in H-band. Direct
comparison suggests that the difference in deviation
and dispersion in zonal velocities between 14 July H-
and 16 July K'-band features at fixed latitude is not
primarily attributable to vertical wind shear. This
further suggests that mechanisms other than vertical
wind shear which dominate dispersion in the zonal
winds can cause changes in the magnitude of dis-
persion and deviation in the zonal winds about the
mean zonal wind profile on timescales of hours to
days.
The authors would like to thank Adam Becker for
useful discussions. This research has been supported in
part by the National Science Foundation Science and
Technology Center for Adaptive Optics, managed by
the University of California at Santa Cruz under coop-
erative agreement No. AST 9876783, as well as by NSF
Grant AST-0908575 to the University of California.
The data presented here were obtained at the W.M.
Keck Observatory, which is operated as a scientific part-
nership among the California Institute of Technology,
the University of California and the National Aeronau-
tics and Space Administration. The Observatory was
made possible by the generous financial support of the
W.M. Keck Foundation. The authors extend special
thanks to those of Hawaiian ancestry on whose sacred
mountain we are privileged to be guests. Without their
generous hospitality, none of the observations presented
would have been possible.
14
Fig. 1 Neptune H- (left column) and K'-band (right column) images taken on 14 July 2009 (top two rows) and 16 July
2009 (bottom two rows) towards the beginning (first and third rows) and end (second and fourth rows) of each night.
Images are shown in units of I/F according to the scale shown in the colorbars to the right of the topmost images in each
filter. For comparison, to the left of the 14 July H-band image taken at 11:20 (UT) (top left panel) we show the same image
with overlayed lines of constant latitude (lines are shown at 60◦S, 30◦S, the equator, and 30◦N) and longitude (a line is
shown at the sub-observer longitude). The sub-observer longitudes of navigated images are indicated in the bottom right
corners of panels.
15
Fig. 2 Mean averaged, high-pass filtered H- (panel a) and K'-band (panel b) images. H-band image is averaged from
the full stack of aligned images from 2009 14 July 11:20:21 - 13:47:44 (UT), and K'-band image is averaged from the stack
of aligned images from 2009 16 July 10:54:57 - 14:20:09 (UT). Color dashed lines indicate the derived apparent orbits
of Galatea (green), Larissa (red), and Despina (orange) in each image. In K'-band Proteus (yellow) is also indicated.
These align well with the true traced moon orbits in each image.
16
Fig. 3 Transformed image, taken in H-band on 2009 July 14 12:50:29 (UT).
17
Fig. 4 Transformed averaged images in strips of fixed latitude (52◦S-30◦S) stacked vertically with ascending time. Vertical
axis indicates the exposure time of the first of the five images composing the averaged image. Repeating horizontal red
lines mark the boundaries between successive image strips. Blue dashed lines indicate representative features which are
relatively consistent in brightness and morphology while yellow dashed lines indicate representative features which are
more ephemeral and change morphology more significantly and rapidly.
18
'
1
K
'
'
1
H
'
Fig. 5 Transformed averaged images in strips of fixed latitude -- identified at the top of each column -- stacked vertically
with ascending time. Vertical axis indicates the exposure time of the first of the five images composing the averaged image.
Second (fourth) column shows a zoomed-in image of the feature shown in the first (third) column, with contour lines at 60%
(blue), 70% (green), and 80% (red) the maximum feature intensity, and our derived feature center indicated to illustrate
our feature tracking method. Identifying names are shown above the second and fourth columns in order to easily indicate
these features in Figures 6 and 7.
19
Fig. 6
20
H1
Fig. 6 Derived zonal drift rates (solid lines and dotted lines) from the longitude positions versus times of each feature
tracked in H-band images on 14 July, from 11:20-13:47 (UT), separated into latitude bins indicated at the top of each panel.
Time is measured in minutes relative to 2009 July 14 11:20:20 (UT). Zonal drift rates expected at the latitude of each
feature according to the smooth Voyager wind profile are shown for comparison (dashed lines). Features are separated by
the mean of their absolute residuals in measured longitude position about their derived zonal drift rates: High-σrl (red)
and Low-σrl (blue). The feature identified as 'H1' in Figure 5 is indicated. Significant dispersion about the smooth Voyager
wind profile is observed, even for Low-σrl features.
21
Fig. 7
22
K1
Fig. 7 Same as Figure 6, now for features tracked in K'-band images on 16 July from 10:54-14:20 (UT). Time is measured
in minutes relative to 2009 July 16 10:54:57 (UT). The feature identified as 'K1' in Figure 5 is indicated. Significant
dispersion about the smooth Voyager wind profile is observed in K'-band on 16 July, even for Low-σrl features.
23
Fig. 8 Distribution of the mean of absolute residuals in measured longitude position about the derived zonal drift rates of
features, σrl. Our two classifications are shown separated by a vertical dashed line: Low-σrl (σrl ≤ 0.7 deg) and High-σrl
(σrl > 0.7 deg).
24
∆ Vlon, σ
∆ Vlon
Fig. 9 Derived zonal velocities of 14 July H-band Low-σrl features shown against the smooth Voyager wind profile (black
solid line). The width of uncertainty in the smooth Voyager wind profile is shown with black dashed lines. The length
of time over which each feature was tracked is indicated by the size of the square in which each feature's zonal velocity is
plotted. Square sizes corresponding to the shortest and longest tracking times are shown in the bottom left corner. There
is significant deviation and dispersion in the zonal velocities from the smooth Voyager wind profile, greater than what is
observed on 16 July in K'-band. Our polynomial fit to the zonal velocities vs. latitudes of Low-σrl features is shown with
a red solid line, and the widths of uncertainty of our fit are shown with red dashed lines. Our polynomial fit is listed
in Table 1 along with the smooth Voyager profile fit of Sromovsky et al. (1993), for comparison. Our polynomial fit to
the 14 July H-band feature zonal velocity distribution deviates significantly from the smooth Voyager wind profile, with a
shift toward more positive velocities of 180±50 m/s. A graphical illustration of our two definitions of deviation in the zonal
velocities, ∆Vlon,σ (red "fake" features) and ∆Vlon (black "fake" features), is shown in the top legend. Green lines
illustrate positive values of deviation while yellow lines indicate negative values of deviation, according to the definition
of each quantity.
25
Fig. 10 Same as Figure 9, now showing the derived zonal speeds of 16 July K'-band Low-σrl features against the smooth
Voyager wind profile (black solid line). There is significant deviation and dispersion in zonal velocities from the smooth
Voyager wind profile. 16 July K'-band feature zonal velocities agree well (within 1σ) with the smooth Voyager wind profile
in their polynomial fit (red solid line).
26
Fig. 11 a) cloud-top pressures versus latitudes of features seen in both H- and K'-bands on 14 and 16 July. Features are
shown in separate bins of the cosine of emission angle, µ, by the color and symbol in which they are plotted, as indicated
in the bottom left corner of the panel. b) cloud-top pressures versus µ of features shown in separate bins of latitude by the
color and symbol in which they are plotted, as indicated in the bottom left corner of the panel. Aside from one feature
at 50◦S, equatorial features are uniformly deeper while features in the north are found at higher altitudes. c) the cloud-top
pressures vs. residuals in zonal velocity from the smooth Voyager wind profile, ∆Vlon, of features for which we were able
to make both measurements, from 32◦S to 28◦N (where vertical wind shear is expected to be important) in thin bins of
latitude. Symbol size indicates the length of time over which a feature was tracked, increasing linearly from the shortest
to the longest tracking times indicated in the bottom right corner of the panel. Squares are used to represent features
whose velocities were obtained by tracking positions in 14 July H-band images and triangles are used to represent features
whose velocities were obtained by tracking positions in 16 July K'-band images.
27
Fig. 12 Derived north-south wind speeds Vlat versus mean latitude positions of 14 July H- (panel a) and 16 July K'-band
(panel b) Low-σrl features. Length of time over which features were tracked is indicated by the size of the symbol in
which they are plotted, increasing linearly from the shortest to longest tracking times indicated in each panel.
28
Fig. 13 a) residuals in zonal velocity from the smooth Voyager wind profile, ∆Vlon, for H- (squares) and K'-band
(triangles) Low-σrl features. Length of time over which features were tracked is indicated by the symbol size in which
each feature is plotted, increasing linearly from the shortest to the longest tracking times in each filter, as indicated in
the panel. Latitude bins in which we average the absolute residuals in zonal velocity of features are indicated by shaded
bars. b) the mean absolute residuals in zonal velocity from the smooth Voyager wind profile, mean ∆Vlon, averaged in the
latitude bins indicated, for 14 July H- and 16 July K'-band Low-σrl features. Symbol size indicates the number of features
composing the mean average, increasing linearly with the number of features. H-band features on average show greater
deviation from the smooth Voyager wind profile. c) Differences between the maximum and minimum zonal velocities of
Low-σrl features within each latitude bin (where two or more features are found in each filter), max δVlon.
2
9
Latitude (deg)
-64.3±.7
-48.0±.9
-45.7±.5
-45.5±.3
-44.0±.6
-42.9±.3
-38.0±.8
-35.0±.8
-30.8±.9
-30.6±.7
-30.5±.7
-23.8±.2
-22.7±.6
-22.5±.5
-1.4±.3
-1.1±.5
-0.2±.5
2.8±.8
6.8±.1
14.4±.3
19.1±.4
19.8±.9
29.9±.8
30.0±.5
41.0±.1
dφ/dt (deg/hr)
5.22±0.28
-0.18±0.23
-5.51±1.20
-1.24±0.20
4.12±1.21
-1.07±0.14
1.41±0.16
-0.32±0.55
-1.29±0.10
-0.73±0.37
1.46±0.29
-2.11±0.14
-7.69±1.16
-6.26±0.95
-7.42±2.12
-2.63±0.72
-0.86±0.26
-1.79±0.14
-1.64±0.20
-1.65±0.65
-3.54±0.34
-1.38±0.31
-1.18±0.13
-0.13±0.32
0.50±1.04
Vlon (m/s) ∆Vlon,σ (m/s)
272±15
-15±19
-462±101
-105±17
356±105
-94±13
133±15
-31±54
-133±11
-75±38
151±30
-232±15
-852±129
-694±105
-891±254
-316±87
-103±31
-215±17
-196±24
-191±76
-401±38
-156±35
-123±13
-13±33
45±95
-22±15
-46±19
350±101
-10±17
385±105
-46±13
246±15
121±54
69±11
129±38
357±30
42±15
527±129
369±105
481±254
69±87
283±31
170±17
181±24
153±76
53±38
152±35
90±13
198±33
117±95
dθ/dt (deg/hr)
0.09±0.10
0.24±0.09
-0.67±0.57
-0.53±0.05
-1.30±0.86
-0.03±0.04
0.18±0.12
-0.55±0.17
-0.31±0.08
0.26±0.07
0.09±0.10
-0.47±0.12
-0.79±0.88
0.64±0.27
-1.82±0.72
-0.39±0.43
0.40±0.20
0.28±0.12
1.26±0.13
-0.20±0.40
-0.11±0.28
0.17±0.26
0.50±0.21
0.54±0.33
-1.17±1.14
Vlat (m/s)
11±12
28±10
-79±68
-63±6
-154±103
-4±5
21±14
-66±20
-37±9
31±8
11±12
-56±14
-93±105
75±32
-215±85
-46±51
47±23
33±14
148±15
-23±47
-13±33
20±30
59±24
64±40
-139±135
σrl (deg) ∆t (min) Nmeas
0.46
0.39
0.21
0.25
0.16
0.27
0.25
0.38
0.17
0.61
0.36
0.19
0.20
0.24
0.53
0.19
0.39
0.23
0.35
0.09
0.50
0.25
0.19
0.42
0.38
139
139
33
139
24
139
123
91
139
139
139
139
33
41
41
41
139
139
139
33
139
91
139
139
74
15
15
5
15
4
15
13
9
15
15
15
15
5
6
6
6
15
15
15
5
15
9
15
15
7
Table 2 Table of various values for Low-σrl features tracked in H-band on 14 July. Definitions of most quantities are given throughout the text. ∆t is the length
of time over which each feature was tracked, and Nmeas is the number of images in which each feature was measured. For a description of the quantity ∆Vlon,σ see
footnote 2.
Latitude (deg)
-67.0±.5
-65.7±.4
-63.3±.6
-47.7±.9
-45.8±.9
-38.0±.2
-37.0±.6
-36.9±.1
-30.2±.1
-30.0±.4
-30.0±.5
-29.8±.2
-29.7±.1
-28.8±.1
-24.4±.4
-24.2±.1
-1.2±.4
0.6±.4
1.1±.4
1.1±.3
1.1±.2
3.7±.8
6.7±.5
29.4±.2
29.7±.6
29.9±.1
30.2±.8
39.1±.40
41.4±.7
dφ/dt (deg/hr)
3.50±0.84
2.84±0.89
-0.42±0.84
-0.95±0.05
-3.35±0.32
-2.49±5.34
-3.31±0.84
-2.30±0.23
-2.64±0.67
-3.48±0.16
-3.45±0.37
-2.37±0.11
-1.28±0.10
-3.30±0.73
-2.15±0.29
-2.39±1.94
-5.27±4.85
-5.83±0.64
-4.80±0.26
-2.83±0.68
-2.09±0.30
-2.92±0.22
-3.87±0.17
-0.40±0.49
-2.30±0.34
-2.70±0.44
-3.74±3.50
-2.04±0.13
-1.61±0.78
Vlon (m/s) ∆Vlon,σ (m/s)
164±39
140±44
-22±45
-77±4
-281±27
-235±505
-317±81
-221±22
-274±70
-361±17
-359±38
-247±11
-133±11
-347±76
-235±32
-262±212
-632±582
-700±77
-576±31
-340±82
-251±37
-350±26
-461±21
-41±51
-240±35
-282±45
-388±363
-190±13
-145±70
-95±39
-78±44
69±45
-15±4
169±27
45±505
117±81
20±22
10±70
96±17
93±38
-20±11
81±11
71±76
33±32
8±212
223±582
290±77
166±31
46±82
135±37
33±26
59±21
177±51
-25±35
15±45
125±363
12±13
-11±70
dθ/dt (deg/hr)
0.63±0.36
0.03±0.31
-0.98±0.44
0.08±0.04
-0.61±0.05
0.26±0.42
1.52±0.37
1.07±0.11
0.20±0.59
-0.1±0.15
0.42±0.15
0.15±0.07
0.32±0.05
-0.1±0.14
-0.86±0.24
0.02±0.15
0.95±0.28
-0.60±0.52
-0.17±0.17
0.07±0.42
0.10±0.62
0.48±0.12
0.12±0.06
0.32±0.28
0.49±0.09
0.01±0.21
0.51±2.48
0.05±0.15
-0.05±1.03
Vlat (m/s)
76±43
3±38
-117±53
9±5
-72±6
31±50
181±44
127±13
24±70
-12±18
50±18
17±9
39±6
-12±17
-101±29
2±17
112±33
-70±62
-20±20
8±50
12±73
56±14
14±7
38±33
58±11
1±25
61±294
6±17
-6±123
σrl (deg) ∆t (min) Nmeas
0.17
0.34
0.16
0.15
0.21
0.14
0.15
0.09
0.05
0.32
0.16
0.37
0.25
0.15
0.20
0.16
0.47
0.12
0.12
0.13
0.05
0.14
0.34
0.11
0.15
0.09
0.33
0.33
0.14
36
67
36
198
89
36
28
67
21
149
67
198
198
36
82
21
23
28
67
36
28
82
149
36
74
36
23
149
36
6
7
6
21
10
6
5
7
4
18
7
21
21
6
9
4
4
5
7
6
5
9
18
6
8
6
4
18
6
Table 3 Table of various values for Low-σrl features tracked in K'-band on 16 July. Definitions of most quantities are given throughout the text. ∆t is the length
of time over which each feature was tracked, and Nmeas is the number of images in which each feature was measured. For a description of the quantity ∆Vlon,σ see
footnote 2.
0
3
31
References
Allison, M., Ferguson, J.W.: Geophys. Res. Lett. 17, 2269
(1990)
Asay-Davis, X.S., Marcus, P.S., Wong, M.H., de Pater, I.:
Icarus 203, 164 (2009)
Baines, K.H., Smith, H.Wm.: Icarus 85, 65 (1990)
Baines, K.H., Mickelson, M.E., Larson, L.E., Ferguson,
D.W.: Icarus 114, 328 (1995)
Borysow, J., Trafton, L., Frommhold, L., Birnbaum, G.:
Astrophys. J. 296, 644 (1985)
Karkoschka, E.: Icarus 215, 759 (2011b)
Karkoschka, E.: Icarus 215, 439 (2011a)
Lecacheux, A.,Zarka, Ph.,Desch, M.D.,Evans, D.R.: Geo-
phys. Res. Lett. 20, 2711 (1993)
Lii, P.S., Wong, M.H., de Pater, I.: Icarus 209, 591 (2010)
Limaye, S.S., Sromovsky, L.A.: J. Geophys. Res. 96, 18941
(1991)
Lindal, G.F., Lyons, J.R., Sweetnam, D.N., Eshleman, V.R.,
Hinson, D.P., Tyler, G.L.: Bull. Am. Astron. Soc. 22,
1106 (1990)
Luszcz-Cook, S.H., de Pater, I., ´Ad´amkovics, M., Ham-
Borysow, J., Frommhold, L., Birnbaum, G.: Astrophys. J.
mel, H.B.: Icarus 208, 938 (2010)
326, 509 (1988)
Borysow, A.: Icarus 92, 273 (1991)
Borysow, A.: Icarus 96, 169 (1992)
Borysow, A.: Icarus 106, 614 (1993)
Colina, L., Bohlin, R.C., Castelli, F.: Astron. J. 112, 307
(1996)
Conrath, B., Flasar, F.M., Hanel, R., Kunde, V., Maguire,
W., Pearl, J., Pirraglia, J., Samuelson, R., Cruik-
shank, D., Horn, L.: Science 246, 1454 (1989)
Conrath, B.J., Flasar, F.M., Gierasch, P.J.: J. Geophys.
Res. 96, 18931 (1991)
Luszcz-Cook, S. 2012, "Millimeter and Near-Infrared Obser-
vations of Neptune's Atmospheric Dynamics", PhD The-
sis, University of California, Berkeley
Martin, S.C., de Pater, I., Marcus, P.: Astrophys. Space Sci.
337, 65 (2012)
Max, C.E., Macintosh, B.A., Gibbard, S.G., Gavel, D.T.,
Roe, H.G., de Pater, I., Ghez, A.M., Acton, D.S., Lai, O.,
Stomski, P., Wizinowich, P.L.: Astron. J. 125, 364 (2003)
Mitchell, J.L., ´Ad´amkovics, M., Caballero, R., Turtle, E.P.:
Nat. Geo. 4, 589 (2011)
Polvani, L.M., Wisdom, J., Dejong, E., Ingersoll, A.P.: Sci-
Conrath, B.J., Gautier, D., Owen, T.C., Samuelson, R.E.:
ence 249, 1393 (1990)
Pravdo, S.H., Shaklan, S.B., Wiktorowicz, S.J., Kulka-
rni, S., Lloyd, J.P., Martinache, F., Tuthill, P.G., Ire-
land, M.J.: Astrophys. J. 649, 389 (2006)
Rages, K., Hammel, H.B., Lockwood, G.W.: Icarus 159,
262 (2002)
Smith, B.A., et al.: Science 246, 1422 (1989)
Sromovsky, L.A., Limaye, S.S., Fry, P.M.: Icarus 105, 140
(1993)
Sromovsky, L.A., Limaye, S.S., Fry, P.M.: Icarus 118, 25
(1995)
Sromovsky, L.A., Fry, P.M., Baines, K.H., Dowling, T.E.:
Icarus 149, 435 (2001a)
Sromovsky, L.A., Fry, P.M., Dowling, T.E., Baines, K.H.,
Limaye, S.S.: Icarus 149, 459 (2001b)
Sromovsky, L.A., Fry, P.M., Dowling, T.E., Baines, K.H.,
Limaye, S.S.: Icarus 150, 244 (2001c)
Sromovsky, L.A., Fry, P.M., Baines, K.H.: Icarus 156, 16
(2002)
Sromovsky, L.A., Fry, P.M., Boudon, V., Campargue, A.,
Nikitin, A.: Icarus 218, 1 (2012)
Stone, E.C., Miner, E.D.: Science 246, 1417 (1989)
van Dam, M.A., Le Mignant, D., Macintosh, B.A.: Proc.
SPIE 5490, 174 (2004)
Warwick, J.W., et al.: Science 246, 1498 (1989)
Icarus 101, 168 (1993)
Crisp, D., Trauger, J., Stapelfeldt, K., Brooke, T., Clarke, J.,
Ballester, G., Evans, R., WFPC2 Science Team: Bull.
Am. Astron. Soc. 26, 1093 (1994)
de Pater, I., Gibbard, S.G., Chiang, E., Hammel, H.B., Mac-
intosh, B., Marchis, F., Martin, S.C., Roe, H.G., Showal-
ter, M.: Icarus 174, 263 (2005)
Fletcher, L.N., Drossart, P., Burgdorf, M., Orton, G.S., En-
crenaz, T.: Astron. Astrophys. 514, A17 (2010)
Gibbard, S.G., Roe, H., de Pater, I., Macintosh, B.,
Icarus
Gavel, D., Max, C.E., Baines, K.H., Ghez, A.:
156, 1 (2002)
Gibbard, S.G., de Pater, I., Roe, H.G., Martin, S., Macin-
tosh, B.A., Max, C.E.: Icarus 166, 359 (2003)
Hammel, H.B.: Icarus 80, 14 (1989)
Hammel, H.B., Baines, K.H., Bergstrahl, J.T.: Icarus 80,
416 (1989a)
Hammel, H.B., Beebe, R.F., de Jong, E.M., Hansen, C.J.,
Howell, C.D., Ingersoll, A.P., Johnson, T.V., Limaye, S.S.,
Magalhaes, J.A., Pollack, J.B., Sromovsky, L.A., Suomi,
V.E., Swift, C.E.: Science 245, 1367 (1989b)
Hammel, H.B., Lockwood, G.J., Mills, J.R., Barnet, C.D.:
Science 268, 1740 (1995)
Hammel, H.B., Lockwood, G.W.: Icarus 129, 466 (1997)
Hammel, H.B., Sitko, M.L., Lynch, D.K., Orton, G.S., Rus-
sell, R.W., Geballe, T.R., de Pater, I.: Astron. J. 134,
637 (2007)
Hansen, J.E., Pollack, J.B.: J. Atmos. Sci. 27, 265 (1970)
Hueso, R., Legarreta, J., Rojas, J.F., Peralta, J., P´erez-
Hoyos, S., del R´ıo-Gaztelurrutia, T., S´anchez-Lavega, A.:
Adv. Spac. Res. 46, 1120 (2010)
Irwin, P.G.J., Teanby, N.A., Davis, G.R., Fletcher, L.N.,
Orton, G.S., Tice, D., Hurley, J., Calcutt, S.B.: Icarus
216, 141 (2011)
Jacobsen, R.A., Owen, Jr., W.M.: Astron. J. 128, 1412
(2004)
Karkoschka, E., Tomasko, M.G.: Icarus 211, 780 (2011)
This manuscript was prepared with the AAS LATEX macros v5.2.
|
1507.02388 | 2 | 1507 | 2015-07-13T03:06:02 | A Venus-Mass Planet Orbiting a Brown Dwarf: Missing Link between Planets and Moons | [
"astro-ph.EP"
] | The co-planarity of solar-system planets led Kant to suggest that they formed from an accretion disk, and the discovery of hundreds of such disks around young stars as well as hundreds of co-planar planetary systems by the Kepler satellite demonstrate that this formation mechanism is extremely widespread. Many moons in the solar system, such as the Galilean moons of Jupiter, also formed out of the accretion disks that coalesced into the giant planets. We report here the discovery of an intermediate system OGLE-2013-BLG-0723LB/Bb composed of a Venus-mass planet orbiting a brown dwarf, which may be viewed either as a scaled down version of a planet plus star or as a scaled up version of a moon plus planet orbiting a star. The latter analogy can be further extended since they orbit in the potential of a larger, stellar body. For ice-rock companions formed in the outer parts of accretion disks, like Uranus and Callisto, the scaled masses and separations of the three types of systems are similar, leading us to suggest that formation processes of companions within accretion disks around stars, brown dwarfs, and planets are similar. | astro-ph.EP | astro-ph |
DRAFT VERSION APRIL 2, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
A VENUS-MASS PLANET ORBITING A BROWN DWARF: MISSING LINK BETWEEN PLANETS AND MOONS
A. UDALSKI1, Y. K. JUNG2, C. HAN2,∗, A. GOULD3, S. KOZŁOWSKI1, J. SKOWRON1, R. POLESKI1,3, I. SOSZY ´NSKI1,
P. PIETRUKOWICZ1, P. MRÓZ1, M. K. SZYMA ´NSKI1, Ł. WYRZYKOWSKI1, K. ULACZYK1, G. PIETRZY ´NSKI1, Y. SHVARTZVALD4,
D. MAOZ4, S. KASPI4, B. S. GAUDI3, K.-H. HWANG2, J.-Y. CHOI2, I.-G. SHIN2, H. PARK2, V. BOZZA5,6
1Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
2Department of Physics, Chungbuk National University, Cheongju 371-763, Republic of Korea
3Department of Astronomy, Ohio State University, 140 W. 18th Ave., Columbus, OH 43210, USA
4School of Physics and Astronomy, Tel-Aviv University, Tel-Aviv 69978, Israel
5Dipartimento di Fisica "E.R. Caianiello", Università di Salerno, Via Giovanni Paolo II 132, 84084, Fisciano (SA), Italy
6Istituto Nazionale di Fisica Nucleare, Sezione di Napoli, Via Cinthia 9, 80126 Napoli, Italy and
∗Corresponding author
Draft version April 2, 2018
ABSTRACT
The co-planarity of solar-system planets led Kant to suggest that they formed from an accretion disk, and
the discovery of hundreds of such disks around young stars as well as hundreds of co-planar planetary systems
by the Kepler satellite demonstrate that this formation mechanism is extremely widespread. Many moons in
the solar system, such as the Galilean moons of Jupiter, also formed out of the accretion disks that coalesced
into the giant planets. We report here the discovery of an intermediate system OGLE-2013-BLG-0723LB/Bb
composed of a Venus-mass planet orbiting a brown dwarf, which may be viewed either as a scaled down version
of a planet plus star or as a scaled up version of a moon plus planet orbiting a star. The latter analogy can be
further extended since they orbit in the potential of a larger, stellar body. For ice-rock companions formed in
the outer parts of accretion disks, like Uranus and Callisto, the scaled masses and separations of the three types
of systems are similar, leading us to suggest that formation processes of companions within accretion disks
around stars, brown dwarfs, and planets are similar.
Subject headings: planetary systems – brown dwarfs – gravitational lensing: micro
1. INTRODUCTION
Brown dwarfs are intermediate in mass between stars
and planets, making them important laboratories for test-
ing theories about both classes of objects, including theo-
ries about atmospheres, binarity, and formation mechanism.
Brown dwarfs are also potentially a laboratory to test the-
ories of planet formation, but detecting planets that orbit
brown dwarfs is challenging. Four super-Jupiter planets or-
biting brown dwarfs have been detected to date by three tech-
niques: direct-imaging (Chauvin et al. 2004; Todorov et al.
2010), radial-velocity (Joergens & Müller 2007), and mi-
crolensing (Han et al. 2013). In the first three cases, the super-
Jupiter planet was within a decade in mass of its host and well
separated from it, suggesting that these systems more resem-
ble scaled down binary stars than they do the extreme mass-
ratio systems that are likely to emerge from accretion disks.
In the fourth case, the separation was substantially closer and
the mass ratio somewhat more pronounced so that the planet
could have formed through the core-accretion scenario that is
thought to be the origin of the solar system's gas giants.
With present technology, substantially lower-mass compan-
ions to brown dwarfs, composed primarily of ice and rock,
can only be discovered using the gravitational microlensing
technique. Microlensing does not rely on light from either the
host or planet, but rather infers the existence and properties of
these bodies from their deflection and magnification of light
from a more distant star that is fortuitously aligned with the
system. As such, the low-luminosity of the brown dwarf and
almost complete absence of light from its low-mass compan-
ion do not interfere in any way with the discovery process.
In this paper, we report the microlensing discovery of an in-
termediate system OGLE-2013-BLG-0723LB/Bb composed
of a Venus-mass planet orbiting a brown dwarf. The system
may be viewed either as a scaled down version of a planet
plus star or as a scaled up version of a moon plus planet orbit-
ing a star, suggesting that formation processes of companions
within accretion disks around stars, brown dwarfs, and planets
may be similar.
2. OBSERVATION
Microlensing event OGLE-2013-BLG-0723 was discov-
ered by the Optical Gravitational Lensing Experiment
(OGLE-IV) (Udalski et al. 2015) on May 12, 2013 (HJD ∼
2456424.5) using its 1.3m Warsaw telescope at the Las Cam-
panas Observatory in Chile. At (α, δ)J2000=(17h34m40s.51,
- 27◦26′53′′
.2) [(l,b) = (- 0◦.02,2◦.83)], the event lies in one
of nine fields toward the Galactic bulge that OGLE observes
& 1 hr- 1, which is sufficient cadence to discover terrestrial
planets.
In fact, the discovery was triggered by a short ∼
1 day, ∼ 40% brightening of the lensing event that was still
in progress at the time of the announcement (Figure 1, left
inset), and which was soon suggested to be of planetary ori-
gin. Hence, when the event underwent a second much larger
(factor 10) anomalous brightening 49 days later, OGLE im-
mediately recognized that it would be crucial to fully charac-
terize this second anomaly. In particular, since such violent
brightening is always caused by the source entering a "caus-
tic" (closed curve of formally infinite magnification – see Fig-
ure 2), a caustic exit was inevitable. Because these caustic
crossings typically last only an hour or so and can occur at
any time during the day, OGLE contacted the Microlensing
Follow Up Network (µFUN) which operates a global network
of narrow-angle telescopes capable of near 24-hour monitor-
ing of individually important events. With the aid of real-time
2
FIG. 1.- Light curve of the microlensing event OGLE-2013-BLG-0723, including models with and without the planet. Left inset shows the planetary anomaly,
which includes not just the obvious spike, but also a more extended low level depression. Right inset shows dense coverage of caustic exit, which permitted
measurement of the Einstein radius θE.
modeling that accurately predicted the time of the exit, the
Wise Observatory 0.46m telescope at Mitzpe Ramon, Israel,
obtained dense coverage of this exit (Figure 1, right inset).
underestimated, the error bars are readjusted by
e = k(e2
0
min)1/2,
+ e2
(1)
where emin is a term used to make the cumulative distribution
function of χ2 as a function of lensing magnification becomes
linear and the other term k is a scaling factor used to make
χ2 per degree of freedom (dof) becomes unity. The former
process is needed in order to ensure that the dispersion of data
points is consistent with error bars of the source brightness
and the latter process is required to ensure that each data is
fairly weighted according to error bars.
The principal data were taken in I band. In addition, OGLE
obtained some V -band data in order to characterize the mi-
crolensed source. Finally, µFUN obtained H-band observa-
tions from the 1.3m SMARTS telescope at Cerro Tololo In-
teramerican Observatory in Chile. These are not used in the
present analysis but will be important for the interpretation of
future follow-up observations to cross-check the lens system
parameters
The light curve of OGLE-2013-BLG-0723 shows a system-
atic decline in the baseline. A similar long term linear trend
(of opposite sign) was seen in OGLE-2013-BLG-0341 and
was eventually traced to a nearby bright star that was gradu-
ally moving toward (in that case) the source star, so that more
of its flux was being "captured" in the tapered aperture used
to estimate the source flux. We searched for such a moving
bright star by examining the difference of two images, from
2004 and 2012. We indeed find a dipole from a bright star
roughly 1.5′′ away, which is the characteristic signature of
such moving stars. Having identified the cause of this trend,
we fit for it and remove it.
Photometry of the event was done using a customized
pipeline based on the Difference Imaging Analysis method
(Alard & Lupton 1998; Wo´zniak 2000). Because the photo-
metric errors estimated by an automatic pipeline, e0, are often
3. LIGHT-CURVE ANALYSIS
The principal features of the light curve are a large double-
horned profile near the peak, which is a signature of binary
microlenses, and a short-lived spike two months earlier, which
is characteristic of planetary companions. Modeling these
features yields the mass ratios and instantaneous locations of
the three bodies relative to the source trajectory. See Figure 2.
However, by themselves, these features do not yield either the
physical or angular scale of the diagram, nor its orientation on
the sky. Rather, the diagram is scaled to the "Einstein radius",
which can be expressed either as an angle θE, or as a physical
length rE projected on the observer plane.
The source star and the Earth's orbit act as "rulers" to de-
termine these two quantities, as well as the angular orienta-
tion. The source angular radius θ∗ is known from its observed
UDALSKI ET AL.
3
TABLE 1
LENSING PARAMETERS
Quantity
χ2/dof
t0 (HJD')
u0
tE (days)
s1
q1
α (rad)
s2
q2 (10- 5)
ψ0 (rad)
ρ (10- 3)
πE,N
πE,E
ds2/dt (yr- 1)
dψ/dt (yr- 1)
IS
IB
Value
4126.8/4073
6484.526 ± 0.037
-0.079 ± 0.002
68.48 ± 0.01
5.07 ± 0.02
3.11 ± 0.02
-1.195 ± 0.003
0.97 ± 0.02
6.61 ± 0.01
-4.936 ± 0.005
1.40 ± 0.02
-0.05 ± 0.01
1.35 ± 0.02
0.81 ± 0.02
-0.50 ± 0.02
19.825 ± 0.006
20.181 ± 0.008
NOTE. - IS and IB represent the baseline magnitudes of the source and
blend, respectively.
FIG. 2.- Geometry of the lensing event. The curve with an arrow rep-
resents the source trajectory. Locations of lens components are marked by
filled dots, where M3 is the planet, M1 is the planet's host, and M2 is the bi-
nary companion. The cuspy closed curves are caustics formed by the lens.
The inset shows an enlarged view around the caustic that produced the short-
term anomaly. The small empty circle on the source trajectory in the inset
represents the source size relative to the caustic. We note that the position
of the planet and the resulting caustic vary in time due to its orbital motion
around its host. Lower/left axes are expressed by angle on the sky in mas,
while upper/right axes show projected size of the system in AU. The direc-
tions North and East on the sky are indicated
temperature and flux. Hence, when it passes over a caustic
(see Figure 2), the finite extent of the source smears out what
would be an extremely sharp peak for a point source (Figure 1,
right inset). Second, Earth's accelerated motion induces dis-
tortions in the light curve, which fix both rE relative the size
of Earth's orbit (1 AU) and the orientation of the diagram rel-
ative to that of Earth's orbit, namely the ecliptic. Hence, we
are able to display both the angular scale and physical orien-
tation in Figure 2. The mass M and distance DL of the lens
are then given by
1
DL
4GM
c2 = rEθE.
+ 1
DS
θE
rE
(2)
=
,
The first equation follows from simple geometry. Once the
distance is known, the second equation follow from Einstein
bending angle α = 4GM/(bc2), where b is the impact parame-
ter of the lens-source approach. See the graphical presentation
of the formalism in Gould (2000).
3.1. Microlens Modeling
Full modeling of the light curve ultimately requires a to-
tal of 19 parameters, including 14 parameters to describe the
lensing system and five flux parameters. Some of these man-
ifest themselves in quite subtle effects. The modeling there-
fore proceeded in several distinct phases. First, we determine
the principal parameters describing the binary-lens system by
removing the data for 25 days around the planetary pertur-
bation and fitting the remaining light curve to seven system
parameters plus four flux parameters (two for each observa-
tory). These nine include three parameters to describe the un-
derlying point-lens event (t0,u0,tE), three to describe the per-
turbations induced by the binary companion (s1,q1, α), one to
describe the ratio of the source radius to the Einstein radius,
ρ = θ∗/θE, and the two-dimensional microlens parallax vector
πE = (πE,N, πE,E).
Here t0 is the time of closest approach to a reference point
in the geometry, which is chosen as the "center of magnifica-
tion" of the caustic traversed by the source, u0 is the impact
parameter in units of θE, and tE is the time required for the
source to cross θE. We choose θE as the Einstein radius of
the mass associated with the caustic, which is the total mass
for the close-binary case, and the mass of the binary compo-
nent located closer to the source trajectory for the wide-binary
case. This convention leads to similar tE for the two cases, e.g.
(Gould et al. 2014).
The three binary parameters are the separation of the com-
ponents in units of θE, the ratio of the companion to primary
component masses, and the angle of the source trajectory rel-
ative to the binary axis at t0 (since parallax causes this to
change with time).
The amplitude of the parallax vector is πE = AU/rE, while
its direction is that of the lens-source relative proper mo-
tion (in the geocentric frame). In contrast to the lens-source
trigonometric relative parallax πrel = πEθE, which is a scalar
quantity, the microlens parallax requires two parameters be-
cause the lens is not directly observed, so the orientation of
the lens-source separation relative to the ecliptic is not known
a priori. See Gould & Horne (2013). Normally, one would not
include parallax in the initial fit, but in this case it is so large
that the projected Einstein radius is substantially smaller than
Earth's orbit, rE ∼ 0.35 AU, so that there is no even approxi-
mately good solution without including πE in the fit.
With this parameterization, we find an excellent fit for the
wide-binary case (s > 1) but only a crude approximation to
the observed light curve for s < 1. We then incorporate two-
parameter orbital motion, essentially the instantaneous rela-
tive transverse velocity of the components in Einstein-radius
units. This refinement results in negligible improvement for
the s > 1 solution, but dramatic improvement in the s < 1
solution, so that the two models provide qualitatively simi-
lar approximations of the data. However, it is found that the
s < 1 case is unlikely due to various reasons as discussed in
Appendix.
We then restore the data from the planetary perturbation and
4
FIG. 3.- Distribution of the microlensing parameters of the best-fit solution, whose central values and errors are shown in Table S1. Color coding indicates
points on the Markov Chain within 1 (red), 2 (yellow), 3 (green), 4 (cyan), 5 (blue) σ of the best fit.
add five additional parameters to the fit: (s2,q2, ψ0, γ), where
γ = (ds2/dt/s2,dψ/dt). Here (s2,q2) are the separation and
mass ratio of the planet relative to its host (the smaller com-
ponent of the binary), ψ0 is the angle between binary axis and
the planet-host axis at t0, and γ is the two dimensional pro-
jected velocity of the planet relative to its host. Table 1 shows
the best fit parameters. In Figure 3, we also present the dis-
tribution of ∆χ2 in the parameter space in order to show the
uncertainties and correlations between the parameters.
3.2. Origin of Parallax and Orbital Signals
The microlens parallax signal is quite strong (∆χ2 > 1100),
but since the amplitude πE = 2.7 is one of the largest ever
detected and hence quite unusual. Therefore, it is prudent to
check whether this signal is coming mainly from one part of
the light curve, which would be suspicious, or from the entire
magnified portion of the light curve, as one would expect.
Figure 4 shows that in fact the signal is coming from the
whole magnified light curve. It also shows that planetary or-
bital signal, while much weaker (∆χ2 = 41), is also broadly
based. As expected, the orbital signal is more compact than
the parallax signal.
3.3. Degeneracy
Because the light curve is well covered, the parameters are
well determined. However, the solution is still subject to three
well known discrete degeneracies. First, there is a degeneracy
between pairs of companion mass ratio and normalized sepa-
ration (q,s), with one lying inside the Einstein radius (s < 1)
FIG. 4.- Cumulative distributions of ∆χ2 from the best-fit (orbit+parallax)
model for the standard (green curve) and parallax (magenta curve) models.
and the other outside. This degeneracy is fully understood
at first (Griest & Safizedah 1998; Dominik 1999) and second
(An 2005) order for the case that the binary is approximated
as static. We find that within this context there is no "close"
UDALSKI ET AL.
5
We find (V -
with the source. That is, in each band, the light-curve flux is
modeled as F(t) = fsA(t) + fb, where A is the magnification.
The source and blend magnitudes are then derived from the
fit parameters fs and fb, respectively.
I,I)0 = (0.52,17.09). From this color and flux,
we derive θ∗ = 0.98 ± 0.08 µas by converting (V -
I) → (V -
K) (Bessell & Brett 1988) and then applying a color/surface
brightness relation (Kervella et al. 2004). The derived color
is well blueward of the overwhelming majority of Galactic
bulge stars, except for horizontal branch stars, which are sub-
stantially brighter. Hence, the source is almost certainly in the
Galactic disk. Disk sources are rare in general (a few percent)
partly because the source must be farther than the lens with
the lenses being mostly in the bulge or fairly far in the disk,
and partly because the column density of bulge stars is higher
than for disk stars. However, given that this particular lens
lies far in the foreground (DL ∼ 0.5 kpc), it is not unexpected
for the source to lie in the disk.
Based on the source color and magnitude, we estimate its
distance DS = (7 ± 1) kpc. The exact distance plays no role
in the determination of θ∗ because this depends only on the
flux F ∝ L/D2
S, not the luminosity L and distance separately,
provided that the source lies behind essentially all the dust.
Given that the source sits 360 pc above the Galactic plane,
this is almost certainly the case. The uncertainty in the source
distance also enters into the uncertainty in the lens distance
through Equation (2). In practice, however, this leads to an
error in source distance of only 4%, which is small compared
to other sources of errors.
Finally, we note that the source color indicates a temper-
ature T ∼ 6100 K, which leads to an estimate of the limb-
darkening coefficient (needed for modeling the caustic exit)
of u = 0.495 (Claret 2000).
In most microlensing events, the blended light fb is just a
nuisance parameter, but in the present case it is potentially
very important. We first note that while the I-band magnitude
is secure, the error in the (V -
I) color (not shown) is quite
large. First, the formal error of the fit is 0.2 mag. In addition,
in contrast to the source flux, which is determined solely from
changes in brightness, the blend flux depends on the accuracy
of the baseline photometry. This can be corrupted by bright
nearby blends, which are plainly visible on the images, and by
stars that are too faint to be identified. Hence, we consider the
color to be only indicative and do not make use of it in further
discussion. However, the I band measurement is too bright to
be the OGLE-2013-BLG-0723LA by about 1.5 mag.
To further understand the origin of this light, we measure
the position of the source relative to the "baseline object",
which is an unresolved blend of the source with the star or
stars giving rise to fb. We find ∆θbase = 55 mas. Since the
blend contains 43% of the baseline light, this implies that it
is naively offset from the source by ∆θblend = ∆θbase/0.43 =
125 mas. However, while the source position can be mea-
sured extremely well because it becomes highly magnified,
the "baseline object" is subject to contamination from unseen
stars as well as the wings of nearby bright stars. We con-
clude that the centroid of blend light is separated from the
source by ∆θblend . 125 mas. Because the surface density of
stars I < 20.1 in this field is . 0.1 arcsec- 2 (Holtzman et al.
1998), the probability that this blended light lies so close to
the source (and lens) by chance is < 1%. It is therefore most
likely a companion to either the source or lens. This issue can
be resolved by obtaining AO images at two epochs to deter-
FIG. 5.- Location of the source star in the color-magnitude of stars in the
nearby field. Also marked are the positions of the centroid of the red clump
giant stars and of the blend. For the first two, the error bars are smaller than
the points. For the last, the magnitude is secure but the error in the color (not
shown) is both large and uncertain, being at least 0.2 mag.
solution, i.e., ∆χ ∼ 2000 between close and wide solutions.
However, if both close and wide binary models are permitted
two degrees of orbital motion, i.e. ds2/dt and dψ/dt, then
the wide solution shows essentially no improvement while the
close solution improves dramatically and matches the wide
solution. In Appendix, we discuss the mathematical and phys-
ical nature of this new variant of the close/wide degeneracy
and show that it implies that the wide solution is strongly fa-
vored. We also present additional corroborating arguments
that this is the case. We therefore consider only the wide so-
lution.
Second, if the source lay on the ecliptic, the entire solution
could be "flipped" relative to the ecliptic without changing
the light curve. Since the source lies just 4◦ from the ecliptic,
this degeneracy is close to exact, with the preferred solution
favored by only ∆χ2 = 38. While this is enough to clearly
favor one solution, the main point is that the two solutions
yield almost identical physical implications.
Finally, Figure 2 shows the small planetary caustics just to
the right of the source trajectory.
In principle, there could
be another solution with these caustics just to the left. How-
ever, this geometry is unable to reproduce the long slow "de-
pression" that follows the planetary anomaly and is due to a
demagnification "valley" that follows the axis defined by the
two small caustics. Hence, the solution is both unique and
measured with good precision.
3.4. Characteristics of the Source Star and Blended Light
We determine the angular source radius θ∗ based on its de-
reddened color (V -
I)0 and brightness I0 (Yoo et al. 2004),
which we determine from their offsets from the centroid of the
red giant clump [(V - I),I]0,clump = (1.06,14.45) (Bensby et al.
2011; Nataf et al. 2013), under the assumption that the source
and the clump giants suffer similar extinction. Figure 5 gives
a graphic representation of this procedure, and also shows the
position of the star (or ensemble of stars) that are blended
6
TABLE 2
PHYSICAL PARAMETERS
Quantity
Mass of the planet
Mass of the host
Mass of the binary companion
Distance to the lens
Projected planet-host separation
Projected separation between binary components
Relative lens-source proper motion (geocentric)
Relative lens-source proper motion (heliocentric)
Value
0.69 ± 0.06 M⊕
0.031 ± 0.003 M⊙
0.097 ± 0.009 M⊙
0.49 ± 0.04 kpc
0.34 ± 0.03 AU
1.74 ± 0.15 AU
3.75 ± 0.33 mas/yr
14.5 ± 1.3 mas/yr
mine whether its proper motion is consistent with the source
or lens. If it proves to be associated with the lens, then this
is a four-body system consisting of a brown dwarf orbited by
a Venus-mass planet, with both orbiting a very low-mass star,
and with this entire group orbiting another star that is roughly
double the mass of the star that contributed to the microlens-
ing event.
3.5. Physical Parameters
The principal physical parameters are given in Table 2.
From the determined masses of the individual lens compo-
nents, it is found that the lens is a triple system that is com-
posed of a Venus-mass planet (0.69 M⊕) orbiting a brown-
dwarf host (0.031 M⊙) that is accompanied by a low-mass
star (0.097 M⊙). The projected planet-host separation is
0.34 AU, while the separation between the binary components
is 1.7 AU. The system is located at a distance 490 pc toward
the Galactic center.
Note that we give the proper motion in both the geocentric
frame in which the measurements were actually made, and the
heliocentric frame, which will be useful for comparison with
future observations. Since DL is known, the conversion from
geocentric to heliocentric frames is straightforward.
3.6. Which Binary Component is Host to the Planet?
From Figure 2 or Table 1, the planet's projected position is
5.07/0.97 = 5.2 times closer than the stellar companion. If the
planet were orbiting the more massive companion, this would
occur with probability < 4%. In addition, in order to ensure
stability of the 3-body system, the brown dwarf would have to
be behind (or in front of) the star-planet system by a at least
factor of 3 larger distance than the projected separation of the
brown dwarf and star. This chance alignment further reduces
the probability by a factor 32. Hence, while this configura-
tion is possible in principle, it has extremely low probability.
A very similar argument applies to a circumbinary configura-
tion, in which the planet's orbit is at least three times larger
than the separation between the brown dwarf and star, but just
happens to be seen in projection close to the brown dwarf.
4. DISCUSSION
OGLE-2013-BLG-0723LBb is a missing link between
planets and moons. That is,
its host OGLE-2013-BLG-
0723LB, being a brown dwarf in orbit around a self-luminous
star, is intermediate between stars and planets, in both size and
hierarchical position. Moreover, the scaled mass and host-
companion separation of OGLE-2013-BLG-0723LB/Bb are
very similar to both planets and moons in the solar system.
See Table 3. Both Uranus and Callisto are believed to have
formed in the cold outer regions of their respective accretion
disks, and are mostly composed of the raw materials of such
regions: ice with some rock. In the case of Uranus, it is be-
lieved to have been formed closer to the current location of
Saturn (10 AU) and to have migrated outward. In the table,
the companion-host separations are scaled to the host mass.
This is appropriate because the "snow line", the inner radius
at which icy solids can form (2.7 AU in the solar system) in-
creases with host mass, probably roughly linearly. A plausi-
ble inference from the first three lines of Table 2 is that these
processes scale all the way from solar-type stars hosting plan-
ets, to brown dwarfs hosting "moon/planets", to giant planets
hosting moons.
However, while OGLE-2013-BLG-0723LBb is almost cer-
tainly orbiting OGLE-2013-BLG-0723LB today, we cannot
guarantee that it was born there. Planets in relatively close
binaries (like OGLE-2013-BLG-0723LA,B) can jump from
one star (or in this case, brown dwarf) to the other, if the
system is dynamically perturbed (Kratter & Perets 2012). If
this is what happened to OGLE-2013-BLG-0723LBb, then
would have been captured from an orbit around OGLE-2013-
BLG-0723LA to its current orbit around OGLE-2013-BLG-
0723LB. It would then be similar to Triton, which was cap-
tured by Neptune out of solar orbit. Table 3 shows that the
Neptune-Triton system in its current configuration is also not
terribly different than a scaled version of OGLE-2013-BLG-
0723LB/Bb.
It will be interesting to model the dynamics of the triple sys-
tem OGLE-2013-BLG-0723LA,B,Bb, but it would be prema-
ture to do so. This is because there is likely a fourth member
of this system that is more massive and luminous than the
other three, and which is separated by of order 100 AU or
less. If confirmed, this member could significantly impact the
system dynamics via the Lidov-Kozai mechanism. The prin-
cipal evidence for this fourth member is that there is excess
light in the aperture that is too bright to be the primary star
OGLE-2013-BLG-0723LA, and is so closely aligned with the
source that it is unlikely to be a random interloper. Because
the source flux is known from the H-band light curve (which
is not included in the present analysis), it will be straightfor-
ward to subtract out the source light from a high-resolution
adaptive optics (AO) image and determine the exact location
of the blended light. If well separated, then its association
with the lens can be established from its proper motion, de-
rived from observations at two epochs, which can be com-
pared with Table 2. If directly superposed in the AO images,
then it is almost certainly associated with either the source or
the lens because of the extremely low probability of a random
interloper. The two possibilities can be distinguished based
on the apparent color because the source is highly reddened
by dust, but the lens sits in front of the great majority of the
dust.
OGLE-2013-BLG-0723LBb is the second terrestrial mass
planet orbiting one member of a binary system composed of
two very low mass objects. The other was OGLE-2013-BLG-
0341LBb. Gould et al. (Gould et al. 2014) showed that if all
low-mass stars were in such systems, roughly one should have
been detected. They noted that no strong statistical inference
could be derived from this fact because, for example, if only
1% of low-mass stars were in such systems, there would be
a 1% chance to find one. Because the question about the fre-
quency of these systems was being framed a posteriori, a 1%
effect is not particularly significant.
However, OGLE-2013-BLG-0723LBb reproduces three of
the four elements that went into their estimate of the fre-
quency of such detections, namely small probability of pass-
UDALSKI ET AL.
7
TABLE 3
COMPANION/HOST PAIRS
Host/Companion
Uranus/Sun
OGLE-2013-BLG-0723LBb/B
Callisto/Jupiter
Triton/Neptune
Mcomp/Mhost
(10- 5)
4.4
6.6
5.7
21
rc- h/Mhost
(AU/M⊙)
19
11
13
46
ing by the planetary caustic, small probability of crossing the
central (binary) caustic, and location in a high-cadence OGLE
fields. The fourth criterion, source magnitude brighter than
IS = 18.5, is not satisfied, since IS ∼ 19.8. If the limit were ex-
tended to IS < 20, the total number of available events would
roughly double. Hence, the prediction would rise to two. With
two predicted and two detected, the probability that these are
extremely rare drops roughly as the square, i.e., roughly 10- 4
probability of this occurring by chance if only 1% of these
stars were in such systems. Hence, while there is still no com-
pelling evidence that these systems are extremely common, at
least they are not extremely rare. Further, the current detec-
tions remain consistent with the hypothesis that these systems
are ubiquitous.
The OGLE project has
received funding from the
National Science Centre,
Poland,
grant MAESTRO
2014/14/A/ST9/00121 to AU. Work by C.H. was supported
by Creative Research Initiative Program (2009-0081561) of
National Research Foundation of Korea. A.G. and B.S.G.
acknowledge support from NSF AST-1103471. B.S.G.,
A.G., and R.W.P. acknowledge support from NASA grant
NNX12AB99G.
APPENDIX
RESOLUTION OF THE CLOSE/WIDE DEGENERACY
In the main text of the paper, we stated that the wide solution was consistent with a static binary while the close solution
required substantial orbital motion to match the wide-solution light curve. At first sight this seems qualitatively reasonable since
close binaries have shorter periods than wide ones. However, only one of these solutions corresponds to the physical system,
while the other is merely a mathematical reflection of it. Hence, the real question is not whether each solution is physically
plausible but what is the relative likelihood where each solution, if it were real, could generate the other as a doppelganger.
Before addressing this, some comments must be in order as why this degeneracy is appearing in this case and how likely it
will be to appear generically. In the simplest cases two solutions appear because to first order, the quadrupole expansion of the
close-binary potential and the tidal expansion of the wide-binary potential are identical and so (to this order) produce identical
caustics (Griest & Safizedah 1998; Dominik 1999). Contemporaneously, it was noticed that in more typical cases, the degeneracy
could remain extremely severe even though the caustics were not actually identical and, indeed, appeared rotated relative to each
other (Fig. 8 of (Afonso et al. 2000)). This was ultimately explained by carrying out the analysis to second order (An 2005). That
is, an additional degree of freedom (basically, rotation) could compensate for the slower convergence of the potential. For wide
binaries, the potential converges more slowly as the mass ratio of the companion increases and as its separation gets closer. This
is well illustrated by small mass-ratio, e.g., planet/star, systems. Topologically, the two caustics are the same, but geometrically
they are very different. The planet has a relatively weak effect on the stellar potential, meaning that the star's caustic ("central
caustic") is well-described by low-order terms and hence subject to the close/wide degeneracy. By contrasts, the planetary caustic
is strongly perturbed and not subject to this degeneracy (unless the separation is very large).
In the present case, the caustic feature in the light curve of the wide solution is associated with the lower-mass companion, and
the mass ratio is relatively large. Hence, one expects that (as in the planetary case) higher order terms will play a relatively large
role. And indeed they prove to be too large to permit the degeneracy analyzed by (An 2005). Orbital motion then provides the
additional degree of freedom to compensate for the non-trivial higher-order terms.
Consider then a wide binary that is qualitatively similar to the present case. Regardless of details, its light curve can always
be mimicked by including the appropriate amount of orbital motion. On the other hand, consider a (real) close binary with the
(s,q) parameters predicted by the wide binary but with so far unspecified orbital parameters. Can this system "generate" a static
wide binary with the same light curve? The answer is: only if its orbital parameters happen to lie in a narrow range around those
predicted by the wide pair. Otherwise, the static wide binary would have been consistent with a wide range of close-binary orbital
motions. Because ∆χ2 is quite large, the fine-tuning requirements of the close solution are severe. This is the principal reason
that we discount it.
However, in addition, if the close solution is spurious and merely tuned to the mathematical requirements of matching the wide
solution, it is likely to have characteristics that are atypical of real systems, which can be a further indication of its spurious
nature. We note two such features. First, the ratio of projected kinetic to potential energy (Dong et al. 2009) is quite low
β ≡(cid:18) Ekin
Epot(cid:19)⊥
=
(sθEDL/AU)3(γ yr)2
M/M⊙
= 0.023
(A1)
compared to typical values 0.1 . β . 0.5. The three physical effects that can produce low β are 1) it is near apocenter of an
extremely eccentric orbit, 2) our line of sight happens to be almost tangent to the orbital velocity vector, or 3) we happen to view
the system edge on, so that its semi-major axis is much larger than its projected separation. Since the unprojected ratio typically
has values Ekin/Epot ∼ 0.5 reducing this number by a factor ∼ 20 via projection effects requires some fine tuning. However, in
the present case there is an additional complication: the third option generates the need for an additional low probability event.
That is, if the two stars are extended along the line of sight, then either this projection must be very great to allow the planet to
have a stable orbit around one star (actually brown dwarf), to the planet itself must be seen in projection next to the pair despite
8
the fact that its orbit is at least three times that of the extended binary. Finally, we note that the source star is four times fainter
in the close model. Since its color (so spectral type) is still the same, this places it twice as far as in the wide solution, where the
density of such stars is about a factor 40 smaller. This is somewhat compensated by the eight times larger phase space, but still
does further reduce the probability of this solution by a factor 5.
All the arguments given above make the close-binary solution very unlikely but they do not exclude it. However, this solution
can also be tested by direct imaging. Since there are compelling reasons to conduct such imaging, possibly at several epochs, no
special effort is required to do this. If these observations reveal a blend that is substantially brighter than the source, then the close
solution will be confirmed. The expected baseline magnitudes of the source and blend are (IS,IB) = (21.28,19.42), respectively.
If the wide solution is correct, we expect the blend and source to be comparable (depending on their I - H colors).
Afonso, C., et al. 2000, ApJ, 532, 340
Alard, C. & Lupton, R. H. 1988, ApJ, 503, 325
An, J. H. 2005, MNRAS, 356, 409
Bensby, T., et al., 2011, A&A, 533, 134
Bessell, M. S.& Brett, J. M. 1988, PASP, 100, 1134
Chauvin, G., et al. 2004, A&A, 425, L2
Claret, A. 2000, A&A, 363, 1081
Dominik, M. 1999, A&A, 349, 108
Dong, S., et al. 2009, ApJ, 695, 442
Gould, A. 2000, ApJ, 542, 785
Gould, A., et al., 2014, Science, 345, 46
Gould, A., & Horne, K. 2013, ApJ, 779, 28
REFERENCES
Griest, K., & Safizedah, N. 1998, ApJ, 500, 37
Han, C., et al., ApJ, 2013, 778, 38
Holtzman, J. A., et al. 1998, AJ, 115, 1946
Joergens, V., & Müller, A. 2007, ApJ, 666, L113
Kervella, P., et al. 2004, A&A, 426, 297
Kratter, K. M. & Perets, H. B. 2012, ApJ, 753, 91
Nataf,D. H., et al., 2013, ApJ, 769, 88
Todorov, K., et al. 2010, ApJ, 714, L84
Udalski, A., Szyma´nski, M.K. & Szyma´nski, G. 2015, Acta Astron., 65, 1
Wo´zniak, P. R. 2000, Acta Astron., 50, 421
Yoo, J., et al., 2004, ApJ, 603, 139
|
1901.02831 | 1 | 1901 | 2019-01-09T17:08:50 | Secular spin-axis dynamics of exoplanets | [
"astro-ph.EP"
] | Context: Seasonal variations and climate stability of a planet are very sensitive to the planet obliquity and its evolution. This is of particular interest for the emergence and sustainability of land-based life, but orbital and rotational parameters of exoplanets are still poorly constrained. Numerical explorations usually realised in this situation are thus in heavy contrast with the uncertain nature of the available data.
Aims: We aim to provide an analytical formulation of the long-term spin-axis dynamics of exoplanets, linking it directly to physical and dynamical parameters, but still giving precise quantitative results if the parameters are well known. Together with bounds for the poorly constrained parameters of exoplanets, this analysis is designed to allow a quick and straightforward exploration of the spin-axis dynamics.
Methods: The long-term orbital solution is decomposed in quasi-periodic series and the spin-axis Hamiltonian is expanded in powers of eccentricity and inclination. Chaotic zones are measured by the resonance overlap criterion. Bounds for the poorly known parameters of exoplanets are obtained from physical grounds (rotational breakup) and dynamical considerations (equipartition of AMD).
Results: This method gives accurate results when the orbital evolution is well known. The chaotic zones for planets of the Solar System can be retrieved in details from simple analytical formulas. For less constrained planetary systems, the maximal extent of the chaotic regions can be computed, requiring only the mass, the semi-major axis and the eccentricity of the planets present in the system. Additionally, some estimated bounds of the precession constant allow to classify which observed exoplanets are necessarily out of major spin-orbit secular resonances (unless the precession rate is affected by the presence of massive satellites). | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. secularspin
January 10, 2019
c(cid:13)ESO 2019
9
1
0
2
n
a
J
9
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
1
3
8
2
0
.
1
0
9
1
:
v
i
X
r
a
Secular spin-axis dynamics of exoplanets
M. Saillenfest, J. Laskar, and G. Boué
IMCCE, Observatoire de Paris, PSL Research University, CNRS, Sorbonne Université, LAL, Université de Lille, 75014
Paris, France. e-mail: [email protected]
Received 28/09/2018; accepted 07/01/2019
ABSTRACT
Context. Seasonal variations and climate stability of a planet are very sensitive to the planet obliquity and its evolution.
This is of particular interest for the emergence and sustainability of land-based life, but orbital and rotational parameters
of exoplanets are still poorly constrained. Numerical explorations usually realised in this situation are thus in heavy
contrast with the uncertain nature of the available data.
Aims. We aim to provide an analytical formulation of the long-term spin-axis dynamics of exoplanets, linking it directly
to physical and dynamical parameters, but still giving precise quantitative results if the parameters are well known.
Together with bounds for the poorly constrained parameters of exoplanets, this analysis is designed to allow a quick
and straightforward exploration of the spin-axis dynamics.
Methods. The long-term orbital solution is decomposed in quasi-periodic series and the spin-axis Hamiltonian is ex-
panded in powers of eccentricity and inclination. Chaotic zones are measured by the resonance overlap criterion. Bounds
for the poorly known parameters of exoplanets are obtained from physical grounds (rotational breakup) and dynamical
considerations (equipartition of AMD).
Results. This method gives accurate results when the orbital evolution is well known. The chaotic zones for planets of
the Solar System can be retrieved in details from simple analytical formulas. For less constrained planetary systems,
the maximal extent of the chaotic regions can be computed, requiring only the mass, the semi-major axis and the
eccentricity of the planets present in the system. Additionally, some estimated bounds of the precession constant allow
to classify which observed exoplanets are necessarily out of major spin-orbit secular resonances (unless the precession
rate is affected by the presence of massive satellites).
1. Introduction
From the works by Laskar & Robutel (1993) and Laskar
et al. (1993b), we know that the long-term dynamics of the
terrestrial planets of the Solar System feature wide chaotic
regions allowing large variations of their obliquity. In par-
ticular, Mars is currently in a chaotic region extending from
0o to 60o obliquity, whereas the Earth is located in a sta-
ble region thanks to the presence of the Moon, resulting in
obliquity variations of only a few degrees. Subsequent stud-
ies detailed both the past and future spin-axis evolution
the Earth (Néron de Surgy & Laskar 1997; Laskar et al.
2004b; Li & Batygin 2014b), Venus (Correia et al. 2003;
Correia & Laskar 2003) and Mars (Laskar et al. 2004a). On
the other hand, paleorecords on Earth show that even the
very slight variations of its orbit and obliquity led to ma-
jor climate changes (e.g. Weertman 1976; Hays et al. 1976).
This implies that life on Earth would be very different to
what it is now if the Earth had evolved in a large chaotic
zone, as it would have without the stabilising effect of the
Moon. This conclusion is reinforced by the fact that high-
obliquity planets undergo severe seasonal variations (e.g.
Spiegel et al. 2009) even for a stable obliquity (but condi-
tions suitable for the very emergence of life could still be
achieved: in some extreme cases, the amount of liquid wa-
ter on the surface may even be favoured by large obliquity
variations, as reported by Armstrong et al. 2014).
As the formation of the Moon is thought to have re-
sulted from an accidental collision event (e.g. Hartmann &
Davis 1975; Canup & Asphaug 2001; Lock et al. 2018), a
"moonless" Earth was a possible (or even likely) outcome
of the planetary formation process. In the broader context
of exoplanets, the dynamics of such a moonless Earth is
thus more than of academic interest. This motivated fur-
ther works about the structure of the chaotic region (Li
& Batygin 2014a) and some additional numerical studies
(Lissauer et al. 2012).
When it comes to exoplanets, we must face the problem
of the incomplete and imprecise nature of both dynami-
cal and physical data. Except from very favourable cases,
like a fast precession and an important flattening induc-
ing detectable transit depth modulations (Carter & Winn
2010; Correia 2014), the spin orientation and the flattening
of exoplanets are far from being reachable by observations.
They must therefore be taken as completely free parame-
ters, in the spirit of the work by Laskar & Robutel 1993
for the Solar System planets. Moreover, the orbital proper-
ties of exoplanets are not well known either, especially the
respective orientations of the orbits (including the mutual
inclinations, which play a crucial role in the spin-axis dy-
namics). Several authors tackled this problem already (e.g.
Brasser et al. 2014; Deitrick et al. 2018; Shan & Li 2018):
they used numerical integrations of the planetary system in
order to build the time-dependent perturbation of the spin
axis. Applied to extrasolar planets, this method requires to
choose a nominal value for both the rotational parameters
and the unknown orbital elements. The parameter space to
be explored is thus very wide, so that even elaborate nu-
merical explorations require some degree of arbitrariness.
Consequently, the use of numerical integrations at this stage
Article number, page 1 of 23
A&A proofs: manuscript no. secularspin
could appear a bit in contradiction with the very incomplete
nature of the data. However, the secular problem (averaged
over rotational and orbital motions) is not as complex as it
could appear. Some hints about a possible analytical treat-
ment were actually given by Laskar (1996) and partially
exploited by Atobe et al. (2004), Li & Batygin (2014a) and
Shan & Li (2018). In the exoplanetary case, a refined ana-
lytical theory would be very convenient, since it would give
in a direct way the sensibility of the spin-axis dynamics
to the various known and unknown parameters, instead of
giving a list of possible outcomes. Associated with some
bounds for the unknown parameters, such a theory would
give a clear range for these possible outcomes.
This was the approach used by Atobe et al. (2004), with
the aim of finding the probability of small obliquity vari-
ations for hypothetical terrestrial planets in the habitable
zone of known exoplanetary systems. Their analytical de-
velopments, though, were limited to the lowest-order ap-
proximations (their parameter space was indeed enormous
since, in this case, the planet itself was hypothetical). The
recent work by Shan & Li (2018) also contains an analytical
part applied to the particular case of exoplanets Kepler-62f
and Kepler-186f. This time, their calculations were mostly
designed to precise and explain numerical results, so they
did not try to bring any substantial improvement to the
theory by Atobe et al. (2004).
In this context, the goal of this article is twofold: i) pro-
vide a general analytical formalism for studying the long-
term spin-axis dynamics of (exo)planets and ii) clarify what
kind of information about the spin can be obtained from
typical observed exoplanetary systems, that is, with nu-
merous unknown physical and dynamical parameters. In
particular, it is crucial for future studies to have a simple
way to classify the observed exoplanets according to the
characteristics of their spin dynamics. This would allow to
determine which exoplanets are worth to be studied in more
details (in particular if a complex chaotic spin dynamics is
expected) and which ones have necessarily a very simple
spin dynamics, making unnecessary any further numerical
or analytical study. Such a general analysis will give both
a qualitative view of the system if it is poorly known (in
the continuation of Atobe et al. 2004), and a quantitative
description of the dynamics if it is well known (as an ana-
lytical counterpart of Laskar & Robutel 1993).
This article is organised as follows: Sect. 2 recalls the
Hamiltonian of the secular spin-axis dynamics and shows
how it can be expanded in terms of the orbital motion
parameters. The secular resonances at all orders can then
be isolated and used to delimit the chaotic regions. Then,
Sect. 3 shows how an incomplete set of orbital elements
can still be used to constrain the orbital solution of an ex-
oplanet. Combined with the rotational breakup limit, it
allows to make a preliminary classification of the "non-
resonant" exoplanets, for which no chaos can appear and
the obliquity variations are constrained by an analytical
bound.
2. Analytical model of the long-term spin dynamics
2.1. Development of the Hamiltonian
Let us consider a system composed of a star and several
planets. We study the rotational dynamics of one planet
among them. For now, we consider that this planet is far
Article number, page 2 of 23
from any spin-orbit resonance. Considering only the lowest-
order term of the torque from the star expanded in Legendre
polynomials, the Hamiltonian of rotation averaged over or-
bital and rotational motions is given for instance by Laskar
& Robutel (1993) and detailed by Néron de Surgy & Laskar
(1997). It can be written
X 2
α
2
(cid:0)1 − e(t)2(cid:1)3/2
(cid:112)
1 − X 2(cid:0)
A(t) sin ψ + B(t) cos ψ(cid:1)
−
+ 2XC(t),
H(X,−ψ, t) = −
(1)
where the conjugate coordinates are X (cosine of obliq-
uity) and −ψ (minus the precession angle). The quantity
α is called the "precession constant" (contrary to previous
studies, we prefer to exclude here the eccentricity e appear-
ing in denominator from the definition of α). Following the
derivation proposed by Néron de Surgy & Laskar (1997),
we obtain
α =
3Gm0
2 ωa3
2C − A − B
2C
.
(2)
In this expression, G is the gravitational constant; m0 is
the mass of the star; a is the semi-major axis of the planet
in orbit around the star; ω is its spin angular velocity, and
A (cid:54) B (cid:54) C are its momenta of inertia. The Hamiltonian (1)
depends explicitly on time t through the eccentricity e and
the functions
A(t) =
B(t) =
2(cid:0) q + pC(t)(cid:1)
(cid:112)
2(cid:0) p − q C(t)(cid:1)
(cid:112)
1 − p2 − q2
1 − p2 − q2
,
C(t) = q p − p q,
(3)
in which q = sin(I/2) cos Ω and p = sin(I/2) sin Ω, where
I and Ω are respectively the orbital inclination and the
longitude of ascending node of the planet. In the following,
we will write η ≡ sin(I/2). One can note that if there is only
one planet in the system (two-body problem), the obliquity
is constant and the precession angle circulates with constant
angular velocity αX/(1 − e2)3/2.
Let us suppose that the eccentricity and the inclination
of the planet are small, such that we can develop the Hamil-
tonian in series of e and η. In the following, we present the
terms up to order 3, but the method presented here can
be generalised to any order (as we will see, the third order
is the first one at which the eccentricity begins to play a
substantial role). Using the fact that C = η2 Ω = O(η2), we
obtain
A = (2 + p2 + q2) q + 2pC + O(η4)
B = (2 + p2 + q2) p − 2q C + O(η4) .
Let us now suppose that the secular orbital dynamics of the
planet, resulting from the perturbations by the other plan-
ets, is quasi-periodic. This amounts to considering that the
chaos present in the orbital secular system acts on a much
larger timescale than the spin dynamics under study. As
we will see below, this holds very well for the Solar System
(this methodology was first proposed by Laskar 1996; it is
(4)
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
used for instance by Li & Batygin 2014a). In this case, we
can write
N(cid:88)
M(cid:88)
j=1
e exp(i) =
η exp(iΩ) =
Ej exp(iθj)
Sj exp(iφj) ,
j=1
where is the longitude of pericentre of the planet in orbit
around the star. The angles θj and φj evolve linearly with
frequencies µj and νj, that is,
j
j
,
(6)
and
φj(t) = νj t + φ(0)
θj(t) = µj t + θ(0)
whereas the amplitudes Ej and Sj are real constants of
order e and η or smaller. Such series can be obtained ei-
ther from analytical theories or from frequency analysis of
numerical solutions (Laskar 1988, 1990). In a general in-
tegrable case, µj and νj are integer combinations of the
fundamental frequencies of the orbital dynamics (usually
noted gk and sk), and the series contain an infinite number
of terms. Arranging the terms by decreasing amplitude, we
consider here a truncation with N terms for the eccentricity
and M terms for the inclination. We get then
εH1(X,−ψ, φ) = −2
1 − X 2
νjSj cos(φj + ψ) ,
(12)
(5)
ε2H2(X, θ, φ) = −
M(cid:88)
j=1
νjS2
j
E2
j + 2X
(cid:112)
M(cid:88)
j=1
3α
4
j=1
X 2
N(cid:88)
N(cid:88)
M(cid:88)
j<k
X 2
j<k
3α
2
−
+ 2X
EjEk cos(θj − θk)
(13)
(νj + νk)SjSk cos(φj − φk) ,
(cid:33)(cid:35)
νkS2
k
cos(φj + ψ)
(cid:32) M(cid:88)
k=1
νjS3
j − 2Sj
νkS2
j Sk cos(2φj − φk + ψ)
e2 =
E2
j + 2
EjEk cos(θj − θk) ,
νiSiSjSk cos(−φi + φj + φk + ψ) .
1 − X 2
and
(cid:34)
(cid:112)
M(cid:88)
ε3H3(X,−ψ, φ) =
−
(cid:112)
M(cid:88)
M(cid:88)
(cid:112)
M(cid:88)
M(cid:88)
1 − X 2
k=1
k(cid:54)=j
j=1
j=1
+
+ 2
1 − X 2
i=1
j<k
j,k(cid:54)=i
− 2
νiSiSjSk cos(−φi + φj + φk + ψ)
+ O(η4) .
In order to obtain an autonomous Hamiltonian, let us in-
troduce the momenta Θj and Φj conjugate to θj and φj.
The system has now N + M + 1 degrees of freedom. The
new Hamiltonian (that we still denote H) can be written
(10)
H = H0 + εH1 + ε2H2 + ε3H3 + O(ε4) ,
where we suppose that O(e) = O(η) = O(ε). The different
M(cid:88)
parts are respectively
N(cid:88)
H0(X, Θ, Φ) = −
α
2
X 2 +
µjΘj +
νjΦj ,
j=1
j=1
F(Σ, σ) = −
with
a = 1 +
3
2
E2
j
1
2
N(cid:88)
M(cid:88)
j=1
j=1
b = νp − 2
νjS2
j
(11)
c = −2 νpSp − νpS3
p + 2Sp
(17)
M(cid:88)
j=1
νjS2
j .
Article number, page 3 of 23
C =
νjS2
j +
j=1
j<k
(νj + νk)SjSk cos(φj − φk) ,
N(cid:88)
M(cid:88)
j=1
N(cid:88)
M(cid:88)
j<k
A sin ψ + B cos ψ = 2
M(cid:88)
(cid:32) M(cid:88)
j=1
k=1
j − 2Sj
and from (4),
+
−
νjS3
j=1
(cid:34)
M(cid:88)
M(cid:88)
M(cid:88)
M(cid:88)
M(cid:88)
k=1
k(cid:54)=j
j=1
i=1
j<k
j,k(cid:54)=i
νjSj cos(φj + ψ)
(cid:33)(cid:35)
νkS2
k
cos(φj + ψ)
νkS2
j Sk cos(2φj − φk + ψ)
(7)
(8)
(9)
(14)
We note that there can be no resonance among the angles φj
and θj because they come from the quasi-periodic solution
of the orbital dynamics. By definition, they are thus already
"integrated".
2.2. One perturbing term: Colombo's top
From (11-12), we conclude that at lowest-order to the per-
turbation, resonant angles can only be of the form σ =
φj + ψ. Let us consider a single resonance with the term
j = p. We introduce the resonant canonical coordinates by
the linear transformation
(cid:26) σ = ψ + φp
ξ = −φp
(cid:26) Σ = −X
Ξ = −X − Φp .
and
Assuming that the system is far from any other resonance,
the long-term dynamics at first order to the perturbation
is given by averaging the Hamiltonian over all angles but
σ. Dropping the constant terms, we get
(15)
a α Σ2 + b Σ + c
1 − Σ2 cos σ ,
(16)
(cid:112)
A&A proofs: manuscript no. secularspin
As shown in Appendix A, the Hamiltonian has the same
form in the case of a 1:1 spin-orbit resonance, with though
a slightly different expression of the coefficients. This would
thus only shift a bit the position of the secular resonances
considered here. In this case, the tidal damping (respon-
sible for this capture in spin-orbit resonance) is supposed
to act on a much larger timescale than the spin dynamics
studied here, such that our approach still holds. Finally,
applying the modified time dτ = a α dt to (16), we obtain
the following Hamiltonian (that we still denote F):
F(Σ, σ) = −
where
1 − Σ2 cos σ ,
Σ2 + γΣ + β
(cid:112)
(18)
1
2
γ =
b
a α
and β =
c
a α
.
(19)
The dynamical system with Hamiltonian function (18) is
well known. It was thoroughly studied by Henrard & Muri-
gande (1987), who called it "Colombo's top" in memory of
Colombo (1966). In the following, we detail its characteris-
tics of interest here, in terms of the two constant parameters
γ and β.
First of all, it is enough to study the case γ (cid:62) 0 and
β (cid:62) 0 since we get the negative cases by the transformations
Σ → −Σ and σ → σ + π, respectively. When studying the
equilibrium points of the system (Henrard & Murigande
1987, or Appendix B.1 and B.2), we find that the phase
space can have two different geometries according to the
value of γ and β (see Fig. 2 for the phase portraits). The
boundary between these two regions of the parameter space
is the curve
C1 =
γ, β (cid:62) 0 : γ2/3 + β2/3 = 1
.
(20)
(cid:110)
(cid:111)
Below C1 (regions A,B,C of Fig. 1), the dynamical sys-
tem (18) has four equilibrium points, that we denote
(a, b, c, d). The equilibrium points c and d merge along the
curve C1, and disappear above it (region D of Fig. 1). Their
respective positions are
Σa ∈ [0, +1]
Σb ∈ [−1, 0]
Σc ∈ [0, +1]
Σd ∈ [0, +1]
, σa = 0 → elliptic (A,B,C,D)
, σb = π → elliptic (A,B,C,D)
, σc = π → hyperbolic (A,B,C)
, σd = π → elliptic (A,B,C) .
(21)
The values Σa,b,c,d have explicit expressions in terms of γ
and β, as given in Appendix B.1 (they correspond to the
different roots of a quartic equation). Since the resonant
angle ψ + φp is equal to 0 or π for each of these equilibrium
points, they all correspond to configurations where the spin
axis, the normal to the orbit (reduced to its pth harmonic)
and the normal to the reference plane are in the same plane.
As such, they are commonly called "Cassini's states" and
labelled (1, 2, 3, 4) after Peale (1969), corresponding to the
equilibrium points (c, a, b, d).
We note the limiting cases
lim
γ→0
lim
γ→0
(cid:112)
for β (cid:54) 1 :
for β (cid:62) 1 :
Σc = 0
Σa = lim
γ→0
Σd = − lim
γ→0
Σb = 0 ,
Σb =
lim
γ→0
Σa = lim
γ→0
Article number, page 4 of 23
1 − β2
(22)
and
for γ (cid:54) 1 :
for γ (cid:62) 1;
lim
lim
β→0
Σc = γ
Σa = lim
β→0
Σd = − lim
β→0
Σb = 1 .
β→0
lim
β→0
Σa = − lim
β→0
Σb = 1
(23)
We are now interested in the width of the resonance,
that is the interval of Σ enclosed in the separatrix emerg-
ing from the hyperbolic fixed point c and containing the
fixed point a. A pendulum approximation can be obtained
for small values of β (as used by Atobe et al. 2004 or Li &
Batygin 2014a), but this approximation is no longer valid
when β grows. Since an analytical expression can be derived
even in the general case, we will use it here. The computa-
tions (Appendix B.5) lead to the following extreme values
of Σ spanned by the resonance:
(cid:113)
Σ± = 2γ − Σc ± 2
They are defined whenever Σc itself is defined (regions
A,B,C of Fig. 1). At this point, it is important to note
that the coordinates (Σ, σ) are singular at Σ = ±1 since
the problem actually takes place on the sphere (Henrard &
Murigande 1987). We must thus study carefully the mean-
ing of the limits (24) when one of them crosses ±1. This
leads to two other limits in the parameter space, as the
curves
−β2 + β(cid:112)
(cid:110)
γ, β (cid:62) 0 : 8 β2 = 1 − 20γ − 8γ2 + (1 + 8γ)3/2 (cid:111)
(cid:110)
8 β2 = 1 + 20γ − 8γ2 + (1 − 8γ)3/2 (cid:111)
0 (cid:54) γ (cid:54) 1/8 , β (cid:62) 0 :
C2 =
C3 =
1 − Σ2
c .
(24)
(25)
(see Appendix B.3-B.4), delimiting the regions A-B and B-
C of Fig. 1, respectively. We note that C1 and C2 intersect
at (γ; β) = (1 ; 0), C1 and C3 intersect at (1/8 ; 3√3/8), and
C2 and C3 intersect at (0 ; 1/2). Contrary to C1, the bound-
aries C2 and C3 do not correspond to actual bifurcations
of the dynamical system, but only to the limits where the
resonant island contains the north pole of the sphere (re-
gion B) and both poles of the sphere (region C). Hence,
the resonance lies in [Σ−; Σ+] in region A; in [Σ−; +1] in
region B; and in [−1; +1] in region C (see Fig. 2). In region
D, there is no more resonance (the separatrix disappears),
but the obliquity can still vary substantially. In Fig. 1, the
colour shades in region D show the oscillation amplitude of
the obliquity as the trajectory passes through Σ = 1.
When β → 0, the width of the island tends to 0 and all
the level curves in the (σ, Σ) plane tend to be horizontal.
In this limit, the resonance width is small and almost inde-
pendent of γ (pendulum approximation). When γ → 0, the
phase portrait in the (σ, Σ) plane tends to be symmetric
with respect to the line Σ = 0.
From (19), we note that
γ =
νp
α
+ O(ε2) and β = −2Spγ + O(ε2) .
(26)
Remembering that Sp is the amplitude of a given term in
the inclination series (5), this means that the parameter
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
Fig. 1. Parameter space of Colombo's top Hamiltonian (18). There is no separatrix (and thus no resonance) in region D, delimited
by the curve C1 (Eq. 20, thick black line). The curves C2 and C3 (Eq. 25, thin black lines) delimit regions A-B and B-C, respectively.
In region A, the resonance island lies between Σ− and Σ+ (Eq. 24); in region B, it lies between Σ− and +1; in region C, it lies
between −1 and +1 (thus the 180o full width given by the colour scale). The problem becomes unphysical above β = 2γ (blue
line) since it corresponds to an amplitude Sp > 1 in the inclination series.
Fig. 2. Examples of phase portraits for the four regions of Fig. 1. The equilibrium points are labelled as in Eq. (21). The level
curves of the Hamiltonian are drawn with black lines out of the resonance island and with red lines inside the resonance. The
parameters chosen are, from A to D: (γ, β) = (0.4, 0.1); (0.55, 0.15); (0.05, 0.7); (0.8, 0.3).
region above the line β = 2γ in Fig. 1 (that is, Sp = 1)
cannot be reached in this problem.
As a simple rule of thumb, one can consider that β con-
trols the resonance width, and that γ controls the location
of the resonance centre (see the Hamiltonian at Eq. 18).
From Eq. (26), we know that β is proportional to the ampli-
tude Sp. Therefore, the resonance widths are larger in hot
planetary systems, for which the mutual inclinations are
large. This was exploited by Boué & Laskar (2010) in their
scenario for tilting the spin-axis of Uranus. On the contrary,
the secular system cannot produce any obliquity variation if
the mutual inclinations are exactly zero. One must keep in
mind that β depends on the precession constant α (Eq. 2)
as well, so that "small" mutual inclinations do not guaran-
tee that the resonances are thin. For the terrestrial planets
of the Solar System, some values of α produce first-order
resonances larger than 70o, even if the mutual orbital incli-
nations are modest (see Sect. 2.3).
Contrary to β, the magnitude of γ cannot be easily
traced back from the planetary architecture: the first-order
secular spin-orbit resonances of any planet can be located
anywhere between 0o and 180o of obliquity. The large ma-
jority of them actually lie in [0o; 90o] because most of the
frequencies νp are negative (the explanation for this prop-
erty is given in Sect. 3.1).
2.3. Overlap of first-order resonances
Going back to the full Hamiltonian (10), the main chaotic
regions of the system can be estimated as the overlap of
the first-order resonances taken separately (Chirikov's cri-
terion). With the analytical expression of their respective
Article number, page 5 of 23
0π2πσ−101ΣregionAaabcd0π2πσ−101ΣregionBaabcd0π2πσ−101ΣregionCaabcd0π2πσ−101ΣregionDaabA&A proofs: manuscript no. secularspin
widths given in Sect. 2.2, the computation of the overlap-
ping regions is straightforward for any quasi-periodic rep-
resentation (5) of the orbital motion.
When comparing two secular terms, we note that no
chaotic zone can form if at least one of them is in region
D, because there is no separatrix in D. This is well verified
by Poincaré sections. A direct consequence of this property
is that if all the terms of the quasi-periodic series are in
region D, the secular dynamics of the obliquity cannot be
chaotic at first order. Moreover, if all the Si coefficients are
small (low inclination regime), so are the corresponding β
coefficients, resulting in virtually no secular variation of the
obliquity (see the shades of grey in Fig. 1). Actually, a no-
chaos criterion can be obtained even if the amplitudes Si are
not known: we just have to check that γi ≈ νi/α > 1 ∀ i =
1...M. On the contrary, if γi < 1 for at least two i, the
existence of a chaotic region is possible, but not guaranteed.
These properties will be discussed further in Sect. 3.4.
This method can be easily checked in the case of the
Solar System, since the secular spin dynamics of all planets
have been studied in the literature. Moreover, a very ac-
curate quasi-periodic approximation of the orbital dynam-
ics can be obtained, since the properties and initial condi-
tions of the planets are very well known. As an example,
the upper row of Fig. 3 shows the widths and overlaps of
first-order resonances for the terrestrial planets (light and
dark-red regions). This figure was produced by applying the
previous analytical formulas to the orbital series of Laskar
(1990), containing more than 50 terms in both eccentricity
and inclination (see Appendix F). We note that most of the
first-order resonances overlap (there are almost no light-red
regions). This results in wide chaotic zones even if the indi-
vidual amplitudes Si are small. However, the full extent of
the chaotic regions given by the frequency analysis (Fig. 3,
second and third rows) cannot be retrieved by only consid-
ering first-order resonances. The following section is thus
dedicated to second and third-order resonances.
2.4. Higher-order resonances
Outside of first-order resonances, we can use a canonical
change of coordinates close to identity in order to suppress
the angular dependency at first order (as already used in a
similar context by Li & Batygin 2014a). Let us consider an
intermediary Hamiltonian X = εX1, such that the current
coordinates are obtained from the new ones through its flow
at time 1. The Hamiltonian in the new coordinates is then
(27)
H = H0 + ε H1 + ε2 H2 + O(ε3) ,
where
H0 = H0
H1 = H1 + {X1,H0}
H2 = H2 + {X1,H1} +
In these expressions, Poisson's brackets are defined as
1
2{X1,{X1,H0}} .
(28)
{f, g} =
∂g
∂qj −
∂f
∂qj
∂g
∂pj
,
(29)
(cid:18) ∂f
(cid:88)
∂pj
j
(cid:19)
where the pairs (pj, qj) are conjugate variables, pj being the
momentum and qj the coordinate. In order to suppress the
Article number, page 6 of 23
angular dependency at order 1, the Hamiltonian X must
fulfil the homological equation
(30)
H1 = H1 + {X1,H0} = H1 ,
in which H1 is the 0th-order term of the multidimensional
Fourier decomposition of H1 (average of H1 over all an-
gles), which is here equal to zero. By matching the terms
of the Fourier decomposition of H1 and X1 one by one, the
solution to the homological equation is
εX1(X, φ,−ψ) = −2
1 − X 2
(cid:112)
M(cid:88)
j=1
νj Sj
νj + αX
sin(φj + ψ) .
(31)
Injecting this function into the expression of the new Hamil-
tonian (28), we get H1 = 0 as required, and the second order
term H2 is given in Appendix C. The only possible resonant
angles at second order have the form σ = φj + φk + 2ψ. The
width of second and higher order resonances is quite small,
so that their separatrices can be computed assuming that
X is near the exact resonance (pendulum approximation).
Accordingly, the centre and half width of the second-order
resonances are computed in Appendix C and given in the
first line of Table 1.
Outside of both first-order and second-order resonances,
the same method can be used to compute the location and
width of third-order resonances. An intermediary Hamilto-
nian of the form X = εX1 + ε2X2 is used, in which X2 must
satisfy a second homological equation (see Appendix D).
The possible resonant angles, as well as the centre and half
width of all the third-order resonances are given in Table 1.
As before, the same method can be used in the case of syn-
chronous rotation, by adding the term −αr(1 + X)2/2 to
H0 (Appendix A). This would only slightly shift the reso-
nances.
With these values, it is straightforward to compute the
overlap regions of every possible second and third-order
resonances. Resonances of order 2 or 3 with a centre lo-
cated inside a resonance of lower order are not considered
(in other words, low-order resonances are supposed unaf-
fected by higher-order ones). The result is shown in the top
row of Fig. 3 for the inner Solar System (blue and green
zones). We obtain a much better match with the frequency
map analysis, showing the importance of high-order reso-
nances in this context. Indeed, numerous frequencies of the
quasi-periodic representation are quite close to each other,
which implies that the corresponding resonances overlap
massively. Hence, even if the second and third-order res-
onances are thin, most of them are all located one after
another, resulting in large chaotic zones. The chaotic dif-
fusion is though slower for higher-order resonances. In the
real Solar System, in which the secular orbital frequencies
are actually not fixed but vary slowly (Laskar 1990), the dif-
fusion of the obliquity will besides be facilitated by small
modulations of the resonances locations.
We note that third-order resonances include terms mix-
ing both eccentricity and inclination (last line of Table 1).
The existence of these terms shows that the s6 + g5 − g6
resonance, which is known to play an important role in the
future obliquity evolution of the Earth (Laskar et al. 1993a,
2004b), has two different origins: one is a first-order reso-
nance with ν23, and the other is a third-order resonance
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
Fig. 3. Top row: Estimate of the chaotic regions of the spin dynamics as the superposition of secular spin-orbit resonances.
The orbital evolution of each planet is approximated by the synthetic representation of Laskar (1990), as detailed in Appendix F.
Light-red and dark-red regions represent the first-order resonances and their overlaps, respectively (Colombo's top Hamiltonian,
Sect. 2.2); light-blue and dark-blue regions represent the second-order resonances and their overlap (Sect. 2.4); and green regions
represent the overlap of third-order resonances. The non-overlapping third-order resonances are not indicated because they are
very thin and thus unimportant for a global picture of the dynamics. Second row: As a comparison, the system given by Eq. (1)
is integrated numerically with the same orbital model (quasi-periodic decomposition of Laskar 1990), and a frequency map analysis
is performed to locate the chaotic zones. The colour scale goes from black (no chaos), to red (strong chaos). Third row: Same
maps obtained from a more detailed model in which the orbital evolution is directly taken from a numerical integration (adapted
from Laskar & Robutel 1993). In the weakly chaotic zones, the dots are shifted vertically according to the level of chaos; in the
strongly chaotic zones, they are plotted in boldface. Bottom row: Same as top row, but the long-term orbital evolution of each
planet is approximated by the Lagrange-Laplace system (Sect. 3.1). The eight planets of the Solar System are included, with the
initial conditions of Bretagnon (1982).
Article number, page 7 of 23
0102030405060020406080100010203040506002040608010001020304050600204060801000102030405060020406080100precessionconstant(arcsec/yr)obliquity(deg)Mercury0102030405060020406080100obliquity(deg)Venus0102030405060020406080100obliquity(deg)Earth0102030405060020406080100obliquity(deg)Mars01020304050600204060801000102030405060020406080100010203040506002040608010001020304050600204060801000102030405060020406080100precessionconstant(arcsec/yr)obliquity(deg)Mercury0102030405060020406080100obliquity(deg)Venus0102030405060020406080100obliquity(deg)Earth0102030405060020406080100obliquity(deg)Mars01020304050600204060801000102030405060020406080100010203040506002040608010001020304050600204060801000102030405060020406080100precessionconstant(arcsec/yr)obliquity(deg)Mercury0102030405060020406080100obliquity(deg)Venus0102030405060020406080100obliquity(deg)Earth0102030405060020406080100obliquity(deg)Mars01020304050600204060801000102030405060020406080100010203040506002040608010001020304050600204060801000102030405060020406080100precessionconstant(arcsec/yr)obliquity(deg)Mercury0102030405060020406080100obliquity(deg)Venus0102030405060020406080100obliquity(deg)Earth0102030405060020406080100obliquity(deg)Mars0102030405060020406080100A&A proofs: manuscript no. secularspin
between µ1, µ10 and ν12 (the numbering refers to Laskar
et al. 2004b). The amplitude of the first resonance, which
is the one emphasised in the literature, is larger by a fac-
tor thirty. This resonance is present in the synthetic repre-
sentation used in this paper (see the term with frequency
−50.30212 (cid:48)(cid:48)/yr in Tables F.1 to F.4). In the top row of
Fig. 3, it appears as a thin isolated resonance (upper-left
corner of the graphs).
Finally, since tidal dissipations are much more efficient
in decreasing the eccentricity than the inclination, one can
imagine a planet with an initially chaotic obliquity wan-
dering in a mixed-type-resonance overlap region, becoming
frozen out of resonance when the amplitudes Ei decrease
due to tidal dissipation. As shown by Laskar et al. 2012,
the eccentricity amplitudes of all the planetary system can
be damped even if only one planet dissipates energy with
the star. Hence the eccentricity modes of an external planet
can be damped even if it is not itself subject to tidal dissi-
pation.
3. Application to exoplanetary systems
In the previous sections, we saw that in the low-eccentricity
and low-inclination regime, the long-term rotational dy-
namics of planets can be studied very efficiently by a simple
analytical model. However, even if numerous exoplanetary
systems are known nowadays, most of the information re-
quired to characterise the rotation of their planets remains
poorly constrained. This information can be split into two
groups: i) the orbital dynamics (amplitudes and frequen-
cies of the quasi-periodic representation) and ii) the rota-
tion parameters (α coefficient). In this section, we will see
how these quantities can be estimated from physical and
dynamical arguments, even with scarce data.
3.1. The Lagrange-Laplace system
Regarding the long-term orbital dynamics, one can use
nominal orbital elements (either best-fit or assumed ones)
and integrate numerically the equations of motion. The so-
obtained solution can then be used directly (as did for in-
stance Brasser et al. 2014, and Deitrick et al. 2018), or
put in the form of quasi-periodic series and used as shown
above. However, this method puts a heavy contrast between
the very uncertain nature of the orbital elements used and
the refined numerical solution applied. Actually, we will see
that at this level of precision, the Lagrange-Laplace sys-
tem is already a good-enough approximation of the orbital
dynamics, up to moderate eccentricities and inclinations
(and without strong effects coming from mean-motion res-
onances). It was used for the same purpose by Atobe et al.
(2004) in the case of a massless hypothetical terrestrial
planet.
The Lagrange-Laplace system is the lowest-order model
of the long-term orbital dynamics: it uses a development of
the Hamiltonian at second order of the eccentricities and
inclinations, which is itself averaged over the fast angles
(secular model at first order to the mutual perturbations).
Let us write
zk = ek exp(ik)
and
ζk = sin
Ik
2
exp(iΩk) ,
(32)
where the index k = 1, 2...N represents a given planet of
the system. Writing z and ζ the vectors of all zk and ζk, the
Article number, page 8 of 23
equations of motion in the Lagrange-Laplace approximation
are
and
ζ = iB ζ ,
z = iA z
where the real matrices A and B are only functions of the
masses and semi-major axes. They can be retrieved from
the lowest-order terms in eccentricity and inclination of the
orbital Hamiltonian. Organising the planets by increasing
semi-major axes, we get from Laskar & Robutel (1995):
(33)
j−1(cid:88)
k=1
2nj
2nj
(cid:18) ak
(cid:19)
(cid:19)
(cid:18) ak
(cid:18) aj
aj
aj
C2
ak
mk
m0
C3
mk
m0
mk
m0
C2
aj
ak
+ nj
(cid:19)
Ajj = nj
Ajk =
and
N(cid:88)
k=j+1
(cid:18) aj
(cid:19)
ak
mk
m0
aj
ak
C3
if k < j
if k > j
N(cid:88)
k=j+1
− nj
mk
m0
aj
ak
C3
(34)
(cid:18) aj
(cid:19)
ak
if k < j
if k > j
(cid:19)
C3
(cid:18) ak
(cid:18) ak
(cid:19)
(cid:19)
(cid:18) aj
aj
aj
C3
ak
mk
m0
C3
aj
ak
j−1(cid:88)
k=1
mk
m0
mk
m0
nj
nj
Bjj = −nj
Bjk =
(35)
in which n2
j = G(m0 + mj), and the functions C2(α) and
j a3
C3(α) are expressed in terms of the Laplace coefficients b(k)
s :
C2(α) =
C3(α) =
3
8
1
4
α b(0)
3/2(α) −
3/2(α) ,
α b(1)
1
4
(1 + α2) b(1)
3/2(α)
(36)
(see Laskar & Robutel 1995, Laskar et al. 2012 or Murray
& Dermott 1999). The equations of motion for z and ζ are
decoupled and linear, such that the solution can be obtained
by diagonalising the matrices A and B. Its expression can
be taken directly from Laskar et al. (2012): it has the form
of quasi-periodic series (5) as required by our model. The
frequencies gk and sk are the eigenvalues of A and B, and
the amplitude of the term k for the planet j is
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Pjk
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)Qjk
N(cid:88)
N(cid:88)
i=1
i=1
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)(cid:12)(cid:12)
P −1
ki zi(0)
Q−1
ki ζi(0)
E(j)
k =
S(j)
k =
(excentricity series)
(inclination series) ,
(37)
where (P, Q) are the matrices composed of the eigenvectors
of (A, B), the matrices (P −1, Q−1) are their inverses, and
zi(0), ζi(0) are the initial conditions of planet i. In this
case, we note that there is a single term for each proper
frequency of the system; the frequencies µj and νj of the
quasi-periodic representation are thus directly equal to one
of the gk and sk, respectively.
Table 1. Critical angle, location and half width of every second and third-order secular spin-orbit resonance.
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
σ
X0
φj + φk + 2ψ
2φj + φk + 3ψ
2φj − φk + ψ
− 1
2α (νj + νk)
− 1
3α (2νj + νk)
− 1
α (2νj − νk)
−φi + φj + φk + ψ
− 1
α (−νi + νj + νk)
φi + φj + φk + 3ψ
− 1
3α (νi + νj + νk)
(cid:112)
K)
K (for a half width 2
8(1 − X 2
0 )
νj νk
(νj−νk)2 SjSk
0 )3/2 α ν2
j νk
(νj−νk)4 S2
j Sk
− 243
2 (1 − X 2
j +3νj νk−ν2
2(2ν2
k)(νj−νk)3+ν2
α(νj−νk)4
j νk(ν2
k−α2)
S2
j Sk
(cid:112)
(cid:112)
1 − X 2
0
2
1
2
P
α(νi−νj )(νi−νk)(2νi−νj−νk)2 SiSjSk
(cid:0)(νi−νj )2+(νj−νk)2+(νk−νi)2(cid:1)
0
1 − X 2
0 )3/2 α νiνj νk
(2νi−νj−νk)2(2νj−νi−νk)2(2νk−νi−νj )2 SiSjSk
− 486(1 − X 2
(cid:112)
1 − X 2
0
φi + θj − θk + ψ
− 1
α (νi + µj − µk)
− 3X0
νi
µj−µk
SiEjEk
Notes. In the fourth line, the symbol P stands for: 4ν5
νjν2
j νk − 3νjν3
k − 2α2νjνk) + νjνk(νj + νk)3.
k − 8ν2
k − 3ν3
k) + νi(ν4
j + ν4
j ν2
i (νj + νk) + ν3
i (13ν2
j + 13ν2
i − 12ν4
k + 16νjνk) − 2ν2
i (3ν3
j + 3ν3
k + ν2
j νk +
From the conservation of total orbital angular momen-
tum, one of the inclination proper frequencies sk is identi-
cally equal to zero (matrix B has rank deficiency of order 1).
The inclination series have thus M = N − 1 terms, whereas
the eccentricity series have N terms. Moreover, all the incli-
nation proper frequencies are negative (this can be shown
from the Geršgorin circles theorem, see Appendix E).
The result, in terms of chaotic zones for the spin dynam-
ics, is shown in Fig. 3, bottom row. Although the match
with the numerical maps (Fig. 3, second and third rows) is
not as good as when we used the synthetic representation
of the orbital dynamics (Fig. 3, top row), the estimate ob-
tained is still remarkably good considering the uncertainties
of the elements of an exoplanetary system1. Using this ap-
proach with the nominal orbital elements given by Brasser
et al. (2014) and Deitrick et al. (2018), we retrieve analyt-
ically their maps showing the possible obliquity variations
of HD 40307 g and Kepler-62 f, in terms of the locations
and widths of the secular spin-orbit resonances (see Fig. 8
by Brasser et al. 2014 and Figs. 5, 6, 10, 11 by Deitrick
et al. 2018). The differences of oscillation amplitude that
they observe are a natural consequence of the initial posi-
tion of the planet with respect to the resonance centre (see
Fig. 2). This shows that the Lagrange-Laplace system, as-
sociated with the development of the Hamiltonian (Sect. 2),
is enough to obtain the level of detail required for study-
ing the long-term rotation of exoplanets up to moderate
eccentricities and inclinations. The use of a more elaborate
model would add no substantial information, owing to the
large uncertainties of the exoplanetary system under study.
3.2. Maximisation of Ek and Sk
Unfortunately, several orbital elements remain unknown for
most of the observed exoplanetary systems. The unknown
elements usually include the mutual inclinations and the
1 Some subtle dynamical effects are not reproduced by the
Lagrange-Laplace system, like the s6 + g5 − g6 first-order reso-
nance mentioned in Sect. 2.4.
relative longitudes of ascending node. From now on, we
suppose that only the masses, the semi-major axes and the
eccentricities are known for all planets of the system. In
this case, the Lagrange-Laplace matrices A and B can still
be computed since they only depend on the masses and the
semi-major axes. We thus obtain the two sets of frequencies
µj and νj. Because the initial conditions zi(0) and ζi(0) ap-
pearing in Eq. (37) are not fully known, the goal here is to
obtain the maximum possible value of the amplitudes Ek
and Sk according to the available data.
For the eccentricity, it amounts to maximise the modu-
lus of a sum of complex numbers with unknown phase. The
result is thus simply the sum of the moduli:
(cid:104)
(cid:105)
max
E(j)
k
= Pjk
(cid:12)(cid:12)P −1
ki
(cid:12)(cid:12) ,
ei
N(cid:88)
i=1
(38)
using the fact that by definition, zi(0) = ei. The problem
is more complex for the inclinations, since both the am-
plitudes and the phases of the initial conditions ζi(0) are
unknown. It is thus necessary to introduce additional ar-
guments, either from physical or from dynamical grounds.
Guided by statistics on the orbital excitation due to close
encounters, Atobe et al. (2004), while dealing mostly with
systems with a single observed planet, imposed I = e/2
for each of them. This law is also in agreement with sta-
tistical distributions of observed exoplanetary systems (Xie
et al. 2016). In our case, though, the application of this
statistical result as a strict rule for each planet of a multi-
planet system seems a bit simplistic. We will opt here for
the hypothesis by Laskar & Petit (2017) of equipartition of
the Angular Momentum Deficit (AMD) among the secular
degrees of freedom. As they point out, this hypothesis is
motivated both by theoretical arguments on chaotic diffu-
sion in the secular dynamics (Laskar 1994, 2008) and by
the aforementioned correlations in observed distributions.
As shown below, this allows to smooth the statistical law
over all the planets contained in the system. Let us intro-
duce the "coplanar AMD" of a planetary system, that is,
Article number, page 9 of 23
A&A proofs: manuscript no. secularspin
the AMD it would have if it was strictly coplanar:
(cid:16)
N(cid:88)
j=1
Cp =
Λj
1 −
(cid:17)
(cid:113)
1 − e2
j
,
where
Λj =
m0mj
m0 + mj
(cid:113)
G(m0 + mj) aj .
(39)
(40)
Contrary to Laskar & Petit (2017), we use here the Hamil-
tonian decomposition of Laskar & Robutel (1995), where
the integrable part is the Sun-planet two-body problem.
This is also the one chosen when expressing the matrices A
and B of the Lagrange-Laplace system (Eqs. 34-35). The
AMD equipartition hypothesis amounts to considering that
the total AMD of the system,
N(cid:88)
(cid:16)
(cid:113)
1 −
1 − e2
j cos Ij
(cid:17)
,
(41)
(42)
C =
Λj
j=1
is equal to
C = 2 Cp .
Hence, even if the individual orbital inclinations are not
known, the so-obtained value of C gives a bound for them.
For instance, the maximum possible value of the inclination
of the kth planet is given by
cos
max Ik
= max
1 −
Λk
, −1
.
(43)
(cid:35)
(cid:34)
N(cid:88)
i=1
sin
Ii
2
Cp
1 − e2
k
(cid:112)
(cid:12)(cid:12)Q−1
ki
(cid:12)(cid:12) ,
obtained from (37) with unknown Ωi, using the con-
straint (42). This constraint can be rewritten
(45)
(cid:112)
1 − e2
where ηi = sin(Ii/2), ci = 2Λi
whereas the quantity to be maximised can be written
i and Z = Cp,
Y =
N(cid:88)
where bi = (cid:12)(cid:12)Q−1
bi ηi ,
i=1
(cid:12)(cid:12). The coefficients ci and bi are all posi-
(46)
ki
tive, and 0 (cid:54) ηi (cid:54) 1. The constraint (45) forms an hyper-
ellipsoid, whereas the quantity to be maximised (46) forms
an hyperplane. Except from particular cases that we will
dismiss here, there is thus only one solution for the max-
imisation of Y , which corresponds to the tangency of the
plane and the ellipsoid. This implies that the two gradients
are collinear:
∇Z = λ∇Y ⇐⇒ ηi = λ
Article number, page 10 of 23
bi
2ci ∀ i = 1, .., N ,
(47)
(cid:104)
(cid:104)
(cid:105)
(cid:105)
max
S(j)
k
= Qjk
N(cid:88)
i=1
Z =
ci η2
i ,
In our case, we are trying to maximise the quantity
Going back to the original notations, this finally gives
Table 2. Secular representation of the Earth orbital dynamics
given by the Lagrange-Laplace theory.
µi
((cid:48)(cid:48)/yr)
3.7137
18.0043
7.3460
17.3308
5.4615
22.2944
2.7015
0.6333
(×105)
1628
1492
1490
1057
404
247
61
1
Ei max[Ei]
(×105)
1974
1917
3286
2381
600
385
194
2
νi
((cid:48)(cid:48)/yr)
0.0000
−18.7456
−6.5701
−5.2008
−17.6358
−25.7514
−2.9039
−0.6778
(×105)
1377
1222
409
425
226
141
87
65
Si max[Si]
(×105)
2420
1989
3424
1606
929
733
993
895
Notes. The eight planets of the Solar System are included. In
the third column, each amplitude is maximised in the case where
both the mutual inclinations and the longitudes are unknown,
assuming the equipartition of AMD between secular degrees of
freedom (since the Solar System is hierarchically AMD stable,
the AMD of the inner and outer parts were taken separately,
see Laskar & Petit 2017). The initial conditions and physical
parameters are taken from Bretagnon (1982).
where λ > 0 by definition of ηi. We get the value of λ from
the imposed value of Z:
N(cid:88)
i=1
Z =
λ2 b2
i
4ci ⇐⇒ λ2 =
Z(cid:80)N
i=1
.
b2
i
4ci
The maximum of Y with the constraint Z is thus
max(cid:2)Y(cid:3) =
(cid:118)(cid:117)(cid:117)(cid:116)Z
(48)
(49)
.
i=1
b2
i
ci
N(cid:88)
(cid:118)(cid:117)(cid:117)(cid:116)Cp
(cid:104)
(cid:105)
(44)
max
S(j)
k
= Qjk
N(cid:88)
i=1
(cid:112)
(Q−1
ki )2
1 − e2
i
2Λi
.
(50)
However, Eq. (47) does not take into account the condi-
tion that all the ηi are smaller than 1. In practice, if the
value obtained for λ implies that one or several ηi are larger
than 1, we just have to fix them to 1 and use the same reso-
lution method iteratively2 with the remaining ηi (changing
the definition of Z and Y accordingly).
Table 2 shows the comparison of the amplitudes ob-
tained for the Earth with the full Lagrange-Laplace system,
and their maximisation supposing that the mutual inclina-
tions and longitudes of node are unknown. As shown in
Fig. 4, a chaotic map can be obtained using these maxi-
mum values. However, we note that each maximisation is
specific to one single amplitude since it implies a distinct
set of inclination values. Taking all the maximum ampli-
tudes at once as if they formed one single representation
gives thus a large upper bound for the chaotic zones. More-
over, due to the large amplitudes of the series obtained,
the small-width approximation for second and third-order
resonances does not necessarily hold.
2 From the form of the constraint (45), decreasing one ηi to 1
implies that at least one of the remaining ηi should increase;
from the solution (47), this actually means that all the remain-
ing ηi increase.
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
the precession constant: instead, we will look for an upper
bound of it from general physical considerations.
After the sphere, the simplest shape model for a ro-
tating planet is given by the Maclaurin ellipsoid (Chan-
drasekhar 1969). It describes the equilibrium shape of a
self-gravitating homogeneous body in rotation with con-
stant angular velocity. The rotational symmetry is imposed
(circular equator), leading to the formula
(cid:112)
(cid:17)
(cid:16)
√1 − 2
,
=
3
(51)
1 − 2
(3 − 22) arcsin − 3
ω2
2πGρ
where ω and ρ are the rotation velocity and the density of
the body and is the eccentricity of its ellipsoidal figure
(in any plane containing the rotation axis). Studying f =
ω2/(2πGρ) as a function of the ellipsoid eccentricity, f is
zero for = 0 and = 1, and it has one maximum at 0 ≈
0.929956 with value f0 ≈ 0.224666. This implies that there
is no such equilibrium figure possible for rotation velocities
larger than
2πGρf0 .
(52)
ωmax =
Converting the ellipsoid eccentricity in terms of momenta
of inertia, we obtain the relation
2C − A − B
(53)
2 .
=
(cid:112)
2C
1
2
Injecting it into the expression of the precession con-
stant (2), we obtain the maximum value
αmax =
3Gm0
4a3
2
0√2πGρf0
.
(54)
For rotation velocities close to ωmax, it is known that there
exist equilibrium ellipsoidal figures with three unequal axes
that have a lower total energy, called Jacobi ellipsoids
(Chandrasekhar 1969). However, we only need an order of
magnitude for αmax and the homogeneous approximation
is anyway quite crude, allowing us to stick to the expres-
sion (54). Planets are expected to spin much more slowly
than ωmax (Eq. 52), including giant gaseous planets (Baty-
gin 2018). In the remaining part of the article, we allow us
to abusively refer to the "rotational breakup" velocity.
Using the average density of the Earth, we obtain a
minimum rotation period of about 2.4 hours, leading to a
maximum precession constant of about 230 (cid:48)(cid:48)/yr. The true
value for the Earth is 20 (cid:48)(cid:48)/yr, or 50 (cid:48)(cid:48)/yr if we include the
additional effects of the Moon (Laskar & Robutel 1993).
This remains well below our bound, but the difference with
this "effective" precession constant due to the presence of
satellites actually constitutes the largest source of uncer-
tainty. Close satellites increase the effective flattening of the
planet, whereas far satellites increase the effective torque
from the star (Boué & Laskar 2006). In both cases, this
increases the precession constant to be used in our model.
This effect is particularly problematic for Saturn, because
our upper bound gives αmax ≈ 0.75 (cid:48)(cid:48)/yr, while the true
value is 0.20 (cid:48)(cid:48)/yr, but it increases to 0.83 (cid:48)(cid:48)/yr if we take the
satellites into account (Ward & Hamilton 2004). Hence, the
introduction of Saturn's satellites makes the precession con-
stant exceed our upper bound. This problem is unavoidable
for exoplanets, because the observation of satellite systems
is hard and none has been observed so far. When using our
model to study the spin dynamics of exoplanets, we must
Article number, page 11 of 23
Fig. 4. Estimate of the chaotic regions of the spin-axis dynamics
for the Earth, where the long-term orbital dynamics is approx-
imated by the Lagrange-Laplace system. The same colour code
as Fig. 3 is used. On the left, the complete set of initial condi-
tions is used (same as Fig. 3, bottom row). On the right, the
orbital elements (I, , Ω) are supposed unknown for all plan-
ets while the remaining ones are taken from Bretagnon (1982).
Accordingly, the coefficients (Ek, Sk) of the quasi-periodic se-
ries are maximised according to the estimated AMD value (see
Sect. 3.2).
For simple systems as those studied by Brasser et al.
(2014), Deitrick et al. (2018) or Shan & Li (2018), the res-
onances are thin and well separated. A picture of the reso-
nant regions (which do not overlap, in these cases) is thus
enough to give clear view of the dynamics. Using the max-
imised amplitudes gives the largest possible widths of the
resonances, which, in turns, show the maximum obliquity
variations and their locations. We fully retrieve their re-
sults. In contrast to these simple and ordered dynamics,
Fig. 5 shows as an illustration the maximised chaotic re-
gions for the GJ 3293 system. The available orbital elements
are taken from Astudillo-Defru et al. (2017), who pointed
out that GJ 3293 d is in the habitable zone. In the spin-down
process toward synchronous rotation due to tidal dissipa-
tive effects from the star (thus decreasing the precession
constant), planet d is the most likely to suffer from large
obliquity changes. One must remember, though, that the
chaotic zones are here maximised according to the avail-
able orbital data. We also predict a rich obliquity dynamics
for the Trappist-1 planets, but due to the confirmed strong
effects of mean-motion resonances (see e.g. Quarles et al.
2017), the use of the Lagrange-Laplace model is probably
inadequate in this case. Building an orbital theory specific
to this system would be out the scope of this paper.
3.3. Maximisation of α
In Sects. 3.1 and 3.2, we saw how to obtain a quasi-periodic
approximation of the long-term orbital motion of a planet,
and how to estimate bounds for its coefficients if all the or-
bital parameters are not known. However, in order to study
its long-term spin dynamics, we still lack an estimate of its
precession constant α. Even in the Solar System, the pre-
cession constants of the planets are not very well known.
The estimate of α for an extrasolar planet would require
observations that are very hard to obtain (its rotation pe-
riod and a model of interior), and which would be specific
to one exoplanet. In order to keep the study as general as
possible, we will not try here to obtain a single value for
01020304050600204060801000102030405060020406080100precessionconstant(arcsec/yr)obliquity(deg)knownamplitudesobliquity(deg)maximisedamplitudesA&A proofs: manuscript no. secularspin
Fig. 5. Chaotic regions of the spin-axis dynamics for exoplanets of the GJ 3293 system, maximised with the method presented
in this paper. The colour code is the same as previous figures. The bounds for their precession constants are, from left to right:
αmax = 66, 20, 6.2 and 1.2 deg/yr (corresponding to maximum rotation periods of a few hours). For each exoplanet, the horizontal
line shows the precession constant corresponding to a rotation period equal to the orbital period (obtained from the method of
Sect. 3.3). They are, from left to right: 13, 31, 48 and 123 days.
bounded by αmax, we can deduce that there can be no
substantial chaotic zone if νi/αmax > 1 whatever the fre-
quency νi. Moreover, if the mutual inclinations are small
(as we assume they are), this implies that the obliquity is
almost constant. Our bound for α (Eq. 54) can thus be used
for a preliminary classification of the exoplanets, while the
bounds for the amplitudes (Eqs. 38,50) are required for a
more specific application to one exoplanet (they allow to
constrain both γ and β, see Fig. 1).
Table 3 shows the 94 planets classified strictly non-
resonant with this criterion, using the Lagrange-Laplace
matrix to estimate the frequencies
(Sect. 3.1) and
Eq. (54) as a bound of α. All the exoplanets from
http://exoplanet.eu with known mass, semi-major axis
and eccentricity were analysed (taking m sin I instead of
the mass if a real-mass estimate was unavailable). At date
2018-03-07, this represents 143 systems with more than one
planet, which contain 353 planets in total (plus the So-
lar System). For some of them, the frequency ratios are so
far from 1 that their classification is quite safe, even when
considering the numerous sources of error inherent to our
method, and in particular, the possible presence of satel-
lites. This mostly concerns planets that are far from their
star, like Uranus and Neptune, and no terrestrial exoplanet
has been observed yet is in this category. Such large semi-
major axes (third column of Table 3) imply that most of the
planets listed in Table 3 are also unaffected by orbital and
rotational tidal dissipation resulting from the interaction
with their central star.
Most of the exoplanetary systems known so far con-
tain only two planets. In this case, there can be no chaotic
region anyway coming from the overlap of first-order reso-
nances (Sect. 2.3) because there is only one forced frequency
in inclination (Sect. 3.1). However, this single term allows
to go one step further and compute the variation range
of the obliquity at first order. This is obtained by looking
at the interval of parameters γ and β (Fig. 1) allowed for
the exoplanet. A very simple formula can be derived if the
inclination amplitude S1 (which is the only amplitude of
the decomposition) is small. Indeed, this implies that β is
small as well (Eq. 26), resulting in quite flat level curves for
Colombo's top Hamiltonian (18). The maximum obliquity
variations are achieved around Σ = 0, and at leading order,
Fig. 6. Empirical mass-radius relationship obtained from the ex-
oplanets with known mass and radius (http://exoplanet.eu).
The exoplanets are supposed to be rocky up to 1 Earth mass,
gaseous beyond 200 Earth masses, and of intermediate composi-
tion in between. Three power laws are used: 1/3, 1/2 and −0.06
from left to right, in general accordance with e.g. Seager et al.
(2007) and Weiss et al. (2013).
thus always keep in mind that the presence of numerous or
massive satellites could modify our conclusions for border-
line cases like Saturn.
Moreover, we must assume a density ρ for the exoplan-
ets if their radius has not been measured, which adds even
more uncertainty. If the radius is unknown, an order of mag-
nitude of the density can be estimated through an empirical
law adjusted to the observed mass-radius distribution (see
Fig. 6). The density is anyway not the major source of un-
certainty of our method.
3.4. Classification of non-resonant exoplanets
In Sect. 2.3, we saw that the overlap of first-order secu-
lar resonances, leading to the largest chaotic zones of the
spin dynamics, can be produced only if the ratio νi/α is
smaller than 1 at least for two frequencies νi of the incli-
nation quasi-periodic representation. Assuming that α is
Article number, page 12 of 23
0510152004590135180024680459013518001234560459013518000.20.40.60.811.204590135180precessionconstant(deg/yr)obliquity(deg)planeteobliquity(deg)planetbobliquity(deg)planetdobliquity(deg)planetc0.111010010−210−1110102103104105radius(Earthradii)mass(Earthmasses)M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
they are equal to
2β
γ ≈ 4S1 .
(55)
∆X ≈
This approximation holds very well for small values of S1.
For the exoplanet WASP-81 c, which has the lowest bound
for S1 in Table 3, we obtain a maximum obliquity excursion
of 0.5o. This limit is very close to what is obtained by plot-
ting the level curves of the Hamiltonian (18), and it holds as
long as α and S1 are below their estimated bounds. Such a
good constraint cannot be achieved for every planet in two-
planet systems, though, since our bound on S1 from the
AMD is sometimes not very informative. It is even dramatic
for planets perturbed by a very massive companion: for ex-
ample the maximum inclination amplitude of HD 92788 c is
higher than 1, indicating that all inclinations are possible.
The maximum excursion of the obliquity is harder to
obtain if there are more than two planets in the system,
since the dynamics is ruled by the superposition of several
forcing terms. However, if the maximum maximised ampli-
tude is small (last column of Table 3), the superposition of
all the terms is unlikely to bring the obliquity over the limit
given at Eq. (55). It can thus be used as well as an order-of-
magnitude estimate. This results in a maximum of about
20o for Uranus and Neptune, showing the crude nature of
our maximisation (as shown by previous works, it is very
hard to tilt Uranus and Neptune by the mean of planetary
perturbations, see e.g. Boué & Laskar 2010).
4. Conclusion
The spin-axis dynamics of a planet plays a major role in
its climate setting, and, by extension, in its suitability for
life. However, the rotation properties of exoplanets are still
very poorly constrained. In this paper, we presented an an-
alytical formulation of the long-term spin-axis dynamics of
a planet, allowing to link known and unknown parameters
to its obliquity evolution and to provide a global picture of
the dynamics in a straightforward way.
At first, the orbital solution is modelled by quasi-
periodic series. This method is thus valid as long as the
orbital chaos, if any, takes place on a much larger timescale
than the spin-axis evolution. The spin-axis Hamiltonian is
then expanded in powers of the eccentricity and inclination
amplitudes of the orbital series. We provided all terms up
to order 3 but the development can be conducted to higher
orders.
A clear picture of the phase space structure is given
by the obliquity ranges associated to the various resonant
regions. The resonant dynamics at order 1 can be charac-
terised analytically in terms of two parameters, which are
linked to the precession constant α (gathering the physical
characteristics of the planet under study) and to the quasi-
periodic representation of the orbit. The pendulum approx-
imation is only used at order 2 and beyond, for which the
resonances are thin enough. The regions of resonance over-
lap at all orders are identified as chaotic. In some cases (as
for the terrestrial planets of the Solar System), these chaotic
regions allow wide excursions of the obliquity. The method
presented here allows to retrieve analytically the previous
numerical results with a good precision. Numerical integra-
tions prove thus to be necessary only if detailed statistics on
the obliquity evolution are required. This is very informa-
tive for Solar System planets (as shown by Néron de Surgy
& Laskar 1997; Correia & Laskar 2003; Laskar et al. 2004a)
but not yet for exoplanets because their initial conditions
and physical parameters are still poorly known. Hence, the
uncertainty of our results remains largely dominated by our
lack of knowledge of the exoplanetary systems rather than
by the approximations inherent to our method. This allows
to stick to the simple analytical formulas presented here.
At this level of uncertainty, the Lagrange-Laplace sys-
tem provides a good-enough representation of the or-
bital motions (excepted for exoplanetary systems featuring
highly excited orbits or strong effects of mean-motion reso-
nances). The formulas obtained allow to set an upper bound
for the amplitude of the eccentricity and inclination terms
if the mutual orientations of the orbits are unknown. On
the other hand, the AMD equipartition hypothesis can be
used, if required, to place a bound on the inclination from
the eccentricity values. Through our analytical model of the
spin-axis dynamics, these maximum amplitudes provide the
maximum extent of the chaotic zones. For example, a large
chaotic region is expected for exoplanet GJ 3293 d for rota-
tional velocities above the synchronous rotation. Systems
very affected by mean-motion resonances (like Trappist-1)
can still be studied using the method described here, but
with the prior construction of a synthetic representation for
the orbital motion, written in the form of a quasi-periodic
series.
However, this method does not allow to consider tidal
dissipations (playing an important role for exoplanets close
to their star), which could be modelled as an adiabatic
process acting on a much longer timescale than the obliq-
uity variations (see Néron de Surgy & Laskar 1997). This
amounts to make the precession constant α and/or the am-
plitudes of the orbital series gradually vary. This method
does not include either the effects of libration around spin-
orbit resonances, even if a trick allows to take into account
a possible locking in synchronous rotation (Appendix A).
Finally, under the hypothesis of hydrostatic equilibrium,
we can set a bound to the precession constant α. This bound
is obtained from the flattening of the planet correspond-
ing to its rotational breakup velocity. Since α governs the
width and location of the resonances, this allows to classify
the exoplanets that cannot be subject to first-order secular
spin-orbit resonances. Among the sufficiently known sys-
tems with more than one planet, we found 94 planets in
this category (26% of our sample). If they belong to ex-
oplanetary systems with low mutual inclinations (as it is
expected in most cases for orbital stability), this implies
that their obliquity is almost constant. This bound for α
is though invalidated by the possible presence of massive
satellites (as our Moon), but some exoplanets are so far
from resonance that their classification is quite safe. This
is the case of Uranus and Neptune.
Considering the high efficiency of the analytical method
proposed here, an obliquity stability map could be designed
easily in the future for each new exoplanet discovered, and
in particular for those classified as "habitable". However,
such a stability map should always be computed again if
any additional planet is found in the system. Indeed, it
would shift the existing frequencies (especially if the new
planet is massive), and add one frequency in both the incli-
nation and eccentricity series, multiplying the possibilities
of resonance. On the other hand, the total AMD of the sys-
tem would increase, resulting in wider maximised chaotic
zones.
Article number, page 13 of 23
Table 3. Exoplanets from http://exoplanet.eu at date 2018-03-07 classified non-resonant with the method detailed in this study.
A&A proofs: manuscript no. secularspin
Name
j/N
HD 113538 b
HD 134987 c
HD 163607 c
HD 89744 c
WASP-53 c
HD 65216 b
HD 85390 c
HD 183263 b
HD 204313 b
HD 27894 d
HD 204313 d
HD 37605 c
HD 45364 b
HD 7449 b
GJ 676A b
TYC+1422-614-1 b
HD 155358 c
HD 34445 g
HD 47366 b
HD 73526 b
HD 37124 d
55 Cnc d
24 Sex b
HD 133131A c
HD 45364 c
HD 102272 c
HD 108874 c
HD 147873 c
HIP 67851 c
HD 125612 d
HD 147018 c
HD 12661 c
HD 74156 c
HD 11506 b
HD 141399 e
HD 4732 c
HD 38529 c
WASP-81 c
HD 154857 c
nu Oph b
24 Sex c
Kepler-419 c
eta Cet b
HD 159868 b
HD 82943 c
mu Ara e
HD 169830 c
1/2
2/2
2/2
1/2
2/2
2/2
2/2
1/2
2/3
3/3
3/3
2/2
1/2
1/2
3/4
1/2
2/2
6/6
1/2
1/2
3/3
5/5
1/2
2/2
2/2
2/2
2/2
2/2
2/2
3/3
2/2
2/2
2/2
2/2
4/4
2/2
2/2
2/2
2/2
1/2
2/2
2/2
1/2
2/2
1/3
4/4
2/2
a
(au)
1.24
5.80
2.42
0.44
3.73
1.30
4.23
1.51
3.17
5.45
3.93
3.81
0.68
2.30
1.81
0.69
1.02
6.36
1.21
0.65
2.81
5.45
1.33
4.36
0.90
1.57
2.68
1.36
3.82
4.20
1.92
2.56
3.82
2.43
5.00
4.60
3.70
2.43
5.36
1.90
2.08
1.68
1.27
2.25
0.75
5.24
3.60
αmax
((cid:48)(cid:48)/yr)
233.0382
4.0960
31.3391
5497.1869
2.0657
242.7438
6.5843
104.8361
9.2689
1.2133
8.4220
5.8515
1672.3835
53.9106
25.8112
1363.7143
647.5095
3.2020
483.1022
1619.2167
30.8760
1.6778
287.8167
9.3548
965.6966
185.6220
34.1534
222.9728
6.7632
3.0777
28.5927
32.4707
4.3343
26.5808
7.2669
7.1381
3.0111
4.6451
4.2372
45.1541
124.2755
52.9030
317.3862
41.0373
391.5783
3.5196
8.7471
min νi
αmax
1.0
1.2
1.2
1.2
1.3
1.4
1.4
1.5
1.5
1.6
1.7
1.8
2.0
2.1
2.2
2.7
2.8
2.8
2.8
2.8
2.9
3.0
3.0
3.4
3.5
3.9
3.9
3.9
4.0
4.1
4.2
4.4
4.7
5.0
5.0
5.4
5.7
5.9
6.6
6.7
6.9
7.0
7.1
7.4
7.9
8.2
8.4
max[Si]
(×104)
1815
763
626
6954
221
596
726
2031
588
268
1343
850
1165
6921
1731
630
754
162
1270
1647
585
34
702
2688
289
3092
885
1300
234
844
357
1213
581
462
1144
730
86
22
898
1183
1300
1024
716
159
1902
405
984
Name
HD 92788 c
HD 113538 c
HD 33844 b
Uranus
HD 47366 c
HD 60532 b
HD 73526 c
HD 128311 b
HD 110014 b
HD 75784 c
HD 89744 b
HD 67087 c
HD 142 c
HD 1605 c
HD 33844 c
47 UMa d
GJ 317 c
ups And d
HD 200964 b
eta Cet c
HD 7449 c
HD 87646A c
HD 82943 b
ups And e
HD 200964 c
HD 183263 c
HD 82943 d
Neptune
HD 5319 b
HD 202206 B
TYC+1422-614-1 c
HD 168443 c
HD 5319 c
HD 30177 b
HD 128311 c
GJ 676 A c
HIP 5158 c
BD+202457 b
NN Ser (AB) d
HD 60532 c
nu Oph c
HD 30177 c
HD 92788 b
BD+202457 c
HIP 57050 c
HD 202206 c
NN Ser (AB) c
j/N
1/2
2/2
1/2
7/8
2/2
1/2
2/2
1/2
2/2
2/2
2/2
2/2
2/2
2/2
2/2
3/3
2/2
3/4
1/2
2/2
2/2
2/2
2/3
4/4
2/2
2/2
3/3
8/8
1/2
1/2
2/2
2/2
2/2
1/2
2/2
4/4
2/2
1/2
1/2
2/2
2/2
2/2
2/2
2/2
2/2
2/2
2/2
a
(au)
0.60
2.44
1.60
19.2
1.85
0.77
1.03
1.10
2.31
6.50
0.88
3.86
6.80
3.52
2.24
11.6
30.0
2.55
1.60
1.93
4.96
1.58
1.19
5.25
1.95
4.25
2.15
30.1
1.75
0.83
1.39
2.84
2.07
3.58
1.76
6.60
7.70
1.45
3.39
1.58
6.10
6.99
0.97
2.01
0.91
2.41
5.38
αmax
((cid:48)(cid:48)/yr)
3698.5423
27.9248
190.8341
0.0682
131.0535
565.4395
406.9428
265.4641
28.8866
1.2344
375.7932
6.1887
0.8689
9.5614
74.3551
0.3274
0.0069
7.8848
162.1684
77.3953
3.7979
17.3656
96.0380
5.6512
137.7501
4.5921
86.0321
0.0156
130.8026
90.1987
72.4459
3.9323
107.4439
4.4350
43.9586
0.5295
0.2293
100.0644
6.7742
39.3544
1.2729
1.0682
115.8082
51.6913
24.6886
9.7781
0.8811
min νi
αmax
8.5
8.6
8.8
9.9
10.5
11.1
11.1
12.2
13.0
16.4
17.2
18.9
20.2
20.7
22.5
23.4
23.9
23.7
28.6
28.9
29.7
31.5
32.1
33.1
33.7
35.0
35.8
43.4
47.8
49.0
51.0
53.8
58.2
63.2
73.5
107.0
134.4
148.1
155.6
159.1
237.0
262.5
272.2
286.7
407.6
452.1
1196.6
max[Si]
(×104)
10293
501
725
1009
967
1291
1309
1563
604
515
1771
2256
317
202
687
569
2182
768
412
446
2617
619
1498
6502
771
1163
10947
810
536
1183
110
401
831
941
643
901
168
682
902
382
587
1808
266
994
24
3521
237
Notes. The first column gives the name of the exoplanet; the second column gives the rank of the exoplanet (sorted by increasing
semi-major axis) and the total number of planets in the system; the third column gives the semi-major axis value; the fourth
column gives the maximum value of the precession constant estimated from Eq. (54); the fifth column gives the minimum ratio of
the eigenfrequencies of the Lagrange-Laplace system and αmax in absolute value (γ parameter of Colombo's top); the sixth column
gives the maximum amplitude of the series decomposition obtained from the maximisation (50), allowing to obtain a maximum
bound for the β parameter of Colombo's top.
Acknowledgements. We thank the anonymous referee for her or his
detailed review.
References
Armstrong, J. C., Barnes, R., Domagal-Goldman, S., et al. 2014, As-
Astudillo-Defru, N., Forveille, T., Bonfils, X., et al. 2017, A&A, 602,
trobiology, 14, 277
A88
Atobe, K., Ida, S., & Ito, T. 2004, Icarus, 168, 223
Batygin, K. 2018, AJ, 155, 178
Boué, G. & Laskar, J. 2006, Icarus, 185, 312
Boué, G. & Laskar, J. 2010, ApJ, 712, L44
Brasser, R., Ida, S., & Kokubo, E. 2014, MNRAS, 440, 3685
Bretagnon, P. 1982, A&A, 114, 278
Article number, page 14 of 23
Canup, R. M. & Asphaug, E. 2001, Nature, 412, 708
Carter, J. A. & Winn, J. N. 2010, ApJ, 716, 850
Chandrasekhar, S. 1969, Ellipsoidal figures of equilibrium (Yale Uni-
versity Press)
Colombo, G. 1966, AJ, 71, 891
Correia, A. C. M. 2014, A&A, 570, L5
Correia, A. C. M. & Laskar, J. 2003, Icarus, 163, 24
Correia, A. C. M., Laskar, J., & de Surgy, O. N. 2003, Icarus, 163, 1
Deitrick, R., Barnes, R., Quinn, T. R., et al. 2018, AJ, 155, 60
Geršgorin, S. 1931, Bulletin de l'Académie des Sciences de l'URSS,
Classe des sciences mathématiques et naturelles, 6, 749
Hartmann, W. K. & Davis, D. R. 1975, Icarus, 24, 504
Hays, J. D., Imbrie, J., & Shackleton, N. J. 1976, Science, 194, 1121
Henrard, J. & Murigande, C. 1987, Celestial Mechanics, 40, 345
Laskar, J. 1988, A&A, 198, 341
Laskar, J. 1990, Icarus, 88, 266
Laskar, J. 1994, A&A, 287, L9
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
Laskar, J. 1996, Celestial Mechanics and Dynamical Astronomy, 64,
Laskar, J. 2008, Icarus, 196, 1
Laskar, J., Boué, G., & Correia, A. C. M. 2012, A&A, 538, A105
Laskar, J., Correia, A. C. M., Gastineau, M., et al. 2004a, Icarus, 170,
Laskar, J., Joutel, F., & Boudin, F. 1993a, A&A, 270, 522
Laskar, J., Joutel, F., & Robutel, P. 1993b, Nature, 361, 615
Laskar, J. & Petit, A. C. 2017, A&A, 605, A72
Laskar, J. & Robutel, P. 1993, Nature, 361, 608
Laskar, J. & Robutel, P. 1995, Celestial Mechanics and Dynamical
Astronomy, 62, 193
Laskar, J., Robutel, P., Joutel, F., et al. 2004b, A&A, 428, 261
Li, G. & Batygin, K. 2014a, ApJ, 790, 69
Li, G. & Batygin, K. 2014b, ApJ, 795, 67
Lissauer, J. J., Barnes, J. W., & Chambers, J. E. 2012, Icarus, 217,
115
343
77
Lock, S. J., Stewart, S. T., Petaev, M. I., et al. 2018, Journal of
Geophysical Research (Planets), 123, 910
Murray, C. D. & Dermott, S. F. 1999, Solar system dynamics (Cam-
bridge University Press)
Néron de Surgy, O. & Laskar, J. 1997, A&A, 318, 975
Peale, S. J. 1969, AJ, 74, 483
Quarles, B., Quintana, E. V., Lopez, E., Schlieder, J. E., & Barclay,
Seager, S., Kuchner, M., Hier-Majumder, C. A., & Militzer, B. 2007,
T. 2017, ApJ, 842, L5
ApJ, 669, 1279
Shan, Y. & Li, G. 2018, AJ, 155, 237
Spiegel, D. S., Menou, K., & Scharf, C. A. 2009, ApJ, 691, 596
Varga, R. S. 2004, Geršgorin and His Circles (Springer-Verlag, Berlin,
Heidelberg)
Ward, W. R. & Hamilton, D. P. 2004, AJ, 128, 2501
Weertman, J. 1976, Nature, 261, 17
Weiss, L. M., Marcy, G. W., Rowe, J. F., et al. 2013, ApJ, 768, 14
Xie, J.-W., Dong, S., Zhu, Z., et al. 2016, Proceedings of the National
Academy of Science, 113, 11431
Article number, page 15 of 23
A&A proofs: manuscript no. secularspin
Appendix A: Case of a 1 : 1 spin-orbit resonance
Numerous exoplanets are observed very close to their star,
in a place where the tidal frictions are strong enough to
efficiently lock them in synchronous rotation. In this sec-
tion, we show that if the librations around the synchronous
rotation are much faster than the secular spin-axis dynam-
ics, we can retrieve Colombo's top Hamiltonian (Sect. 2.2),
allowing to use the same approach as in the non-resonant
case. As before, though, we will not consider the effect of
the tidal dissipation on the obliquity. This is thus only valid
for systems for which the tidal damping of the obliquity acts
on a larger timescale than the spin-axis dynamics.
We will use the same method as Correia et al. (2003).
Let us write λ the mean longitude of the planet in orbit
around the star, and (cid:96) its rotation angle. The mean longi-
tude λ is measured from the equinox at a reference epoch
(for instance J2000), whereas the rotation angle (cid:96) is mea-
sured from the equinox of the date up to a fixed point of the
equator (principal axis A). If we keep the angles of the form
(cid:96) − λ during the average over the mean longitude and the
fast rotation angles (see Néron de Surgy & Laskar 1997),
the corresponding "semi-averaged" Hamiltonian is
L2
2C
H(L, Λ, Y, (cid:96), M,−ψ, t) =
+ nΛ −
(L + Y )2 cos(cid:2)2((cid:96) − λ − ψ)(cid:3)
L2 − Y 2(cid:0)
A(t) sin ψ + B(t) cos ψ(cid:1) + 2Y C(t) ,
(cid:112)
−
−
αr
2L
Y 2
L(cid:0)1 − e(t)2(cid:1)3/2
α
2
(A.1)
where we neglected terms of order e(B − A)/C. The mo-
menta L = Cω and Y = LX are conjugate to (cid:96) and −ψ,
respectively. The momentum Λ, conjugate to λ, has been
added such that λ = n (mean motion). The resonant pre-
cession constant is defined as
αr =
3Gm0
8 ωa3
B − A
C
,
(A.2)
using the same notation as Eq. (2). We note that the angle
λ+ψ appearing in the Hamiltonian corresponds to the mean
longitude measured from the equinox of the date. Let us use
the canonical change of coordinates
(cid:26) θ = (cid:96) − λ
γ = λ
(cid:26) I = L
and
Γ = L + Λ .
(A.3)
Y 2
α
2
αr
2I
The momentum Γ is an arbitrary constant of motion and
the Hamiltonian becomes
H(I, Y, θ,−ψ, t) =
I 2 − Y 2(cid:0)
I 2
2C − nI −
I(cid:0)1 − e(t)2(cid:1)3/2
A(t) sin ψ + B(t) cos ψ(cid:1) + 2Y C(t).
(I + Y )2 cos(2θ − 2ψ)
−
−
We will now suppose that the dynamics of θ, corresponding
to the "semi-secular" timescale (either circulation or oscilla-
tion), is much faster than the evolution of the other degrees
of freedom, corresponding to the secular timescale. We thus
(cid:112)
(A.4)
Article number, page 16 of 23
consider for now that except (I, θ), all the variables are fixed
(adiabatic approximation). The equations of motion are
∂H
I = −
∂θ
∂H
θ =
∂I
=
(I + Y )2
= −αr
I
C − n +
αr
2
I
α
2
I 2 − Y 2
(cid:0)
I
I 2
√I 2 − Y 2
−
−
Y 2
sin(2θ − 2ψ)
I 2(cid:0)1 − e(t)2(cid:1)3/2
A(t) sin ψ + B(t) cos ψ(cid:1) .
cos(2θ − 2ψ)
(A.5)
Using the definition of I, Y and αr, the first equation gives
3Gm0
8a3
B − A
C
(1 + X)2 sin(2θ − 2ψ),
ω = −
resulting, for any value of X, to two equilibrium points:
θ = ψ and ψ+π/2 mod π. We note that θ = ψ is an elliptic
equilibrium while θ = ψ + π/2 is hyperbolic. Injecting this
into the second equation, we obtain
(A.6)
I
θ =
C − n + small terms,
(A.7)
in which the small terms correspond to the precession of
the spin axis (α and αr) and the precession of the orbit (A
and B). The equilibrium condition, corresponding to the
exact resonance, is thus ω ≈ n. Considering that the planet
is locked in synchronous rotation, we have thus θ = ψ and
ω ≈ n. According to the adiabatic approximation, this will
be verified whatever the value of the slow variables, such
that we can inject them into the full Hamiltonian:
X 2
(cid:0)1 − e(t)2(cid:1)3/2 −
α
2
αr
2
(A.8)
(1 + X)2
H(X,−ψ, t) = −
1 − X 2(cid:0)
A(t) sin ψ + B(t) cos ψ(cid:1) + 2XC(t) ,
(cid:112)
−
where this time, we use X as conjugate momentum of −ψ
(the Hamiltonian is thus divided by the constant L). In the
expression of α and αr, we must replace ω by n. We get
here one extra term with respect to (1), due to the spin-
orbit resonance. Using the same method as in Sect. 2.2,
the Hamiltonian in case of a first-order secular spin-orbit
resonance is
1
2
F(Σ, σ) = −
(a α + αr)Σ2 + (b + αr)Σ + c
1 − Σ2 cos σ,
(A.9)
which must be compared to (16). This Hamiltonian has the
same general form and it can be reduced to Colombo's top.
We can thus apply the same method of resolution (redefin-
ing the constants accordingly).
(cid:112)
Appendix B: Characteristic quantities of
Colombo's top
Appendix B.1: Equilibrium points
From (18), the equations of motion are
(cid:112)
Σ = −
σ = +
∂F
∂σ
∂F
∂Σ
= β
1 − Σ2 sin σ
= −Σ + γ − β
Σ
√1 − Σ2
cos σ .
(B.1)
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
(cid:112)
Apart from the coordinate singularity at Σ = ±1, the first
equation implies that Σ = 0 when σ = 0 or π. Injecting this
into the second equation, we get
∆4 = 16γ2β2(cid:104)
1 − Σ2 = ±βΣ
(B.2)
(γ − Σ)
where β (cid:62) 0 by hypothesis. The resolution of this equation
requires to square left and right-hand terms, loosing the
information3 about the sign of cos σ. We obtain a quartic
equation in Σ:
P4(Σ) = Σ4− 2γΣ3 + (γ2 + β2− 1)Σ2 +2γΣ− γ2 = 0 , (B.3)
with discriminant
− γ6 + 3(1 − β2)γ4
− 3(1 + 7β2 + β4)γ2 + (1 − β2)3(cid:105)
(cid:112)
P4(Σ)(cid:12)(cid:12)β=0 = (Σ − 1)(Σ + 1)(Σ − γ)2
(cid:112)
P4(Σ)(cid:12)(cid:12)γ=0 = (Σ −
1 − β2)(Σ +
1 − β2)Σ2 ,
It is zero for the particular cases γ = 0 or β = 0, for which
the polynomial can be factored into, respectively,
(B.4)
.
(B.5)
showing the corresponding solutions and their multiplici-
ties. They constitute equilibrium points of the system when-
ever they are real and in the interval [−1; 1].
For γ > 0 and β > 0, the discriminant can be either
negative (two equilibrium points), zero (three equilibrium
points among which one double root), or positive (four equi-
librium points). The corresponding solutions can be written
analytically according to the general resolution of quartic
equations. They are namely
Σa =
Σb =
Σc =
Σd =
1
2
1
2
1
2
1
2
γ − V +
γ − V −
γ + V −
γ + V +
1
2
1
2
1
2
1
2
2C − D + γ
2C − D + γ
2C − D − γ
2C − D − γ
1 + β2
V
1 + β2
V
1 + β2
V
1 + β2
V
,
(B.6)
where numerous intermediary variables are required in or-
der to get compact expressions:
(cid:114)
(cid:114)
(cid:114)
(cid:114)
W = γ2 + β2 − 1
(cid:112)
Z = 108γ2β2 + 2W 3
(cid:114)
(cid:16)
U = 3
Z +
1
2
Z 2 − 4W 6
(cid:17)
(cid:19)
W
2
3
(cid:18)
C = γ2 −
W 2
U +
D =
U
√C + D .
V =
1
3
1
2
(B.7)
We note that Σc,d are real solutions only when ∆4 (cid:62) 0 (see
below for the limit in terms of γ and β). The corresponding
values of σ are
(B.8)
σa = 0 , σb = π , σc = π , σd = π .
The points a, b and d are elliptic fixed points, whereas the
point c is hyperbolic.
3 After having computed one solution Σ0, this information is
retrieved by checking the sign of Σ0/(γ − Σ0).
Appendix B.2: First boundary (BC/D)
The zero value of (B.4) corresponds to a bifurcation. Its
position can be computed by solving the equation ∆4 = 0,
which corresponds to solving a cubic equation either in γ2
or β2. Choosing to solve it in terms of β, the discriminant
is
(B.9)
∆ = −19683 γ4(1 + γ2)2 < 0 ,
meaning that there is only one real solution. This solution
is
(cid:16)
1 − γ2/3(cid:17)3
γ2 =
β2 =
which is the boundary C1 (20).
or
1 − β2/3(cid:17)3
(cid:16)
,
(B.10)
Appendix B.3: Second boundary (A/B)
The other two boundaries can be obtained by studying the
level curves of the Hamiltonian passing through Σ = ±1
(which is singular using the coordinates Σ and σ, but it
does not matter here).
Let us begin with the +1 case, for which the Hamilto-
nian has value −1/2 + γ. We now look for this specific level
curve along the axes σ = 0 and σ = π. This leads to the
equation
1
2
1
2
+ γ ,
Σ2 + γΣ ± β
1 − Σ2 = −
−
for which Σ = +1 is a solution. By reorganising the terms,
taking the square (thus loosing the information about the
sign of cos σ), and dividing by (Σ − 1), we get
+ γ2 + β2
P3(Σ) =
(cid:18) 1
(cid:19)
(cid:19)
(cid:18)
Σ3 +
Σ2 +
(B.11)
Σ
(cid:112)
4 − γ
1
4
−
(cid:19)
+
+ β2 + γ − γ2
= 0 ,
which is a cubic equation in Σ. Its determinant is
(B.12)
.
(cid:21)
1
4
1
4
−
(cid:18)
(cid:20)
−β4 +
(cid:18) 1
(cid:19)
P3(Σ)(cid:12)(cid:12)β=0 =
P3(Σ)(cid:12)(cid:12)γ=0 =
∆3 = β2
4 − 5γ − 2γ2
β2 + γ(1 − γ)3
(B.13)
Once again, it is zero for β = 0. Moreover the solutions
for γ = 0 can be easily computed. In these two particular
cases, the polynomial can be factored into, respectively,
1
4
1
4
(cid:112)
(Σ − 1)(Σ + 1 − 2γ)2
1 − 4β2)(Σ +
(Σ −
(cid:112)
1 − 4β2)(Σ + 1) ,
(B.14)
showing the solutions and their multiplicities. For γ > 0
and β > 0, the discriminant can be either negative (one so-
lution), zero (three solutions among which one double root),
or positive (three solutions). The zero value corresponds to
the limit we are looking for. Its position can be computed
by solving the equation ∆3 = 0, which amounts to solving a
quadratic equation in β2 or a quartic equation in γ. Choos-
ing to solve it in terms of β, the only positive solution is
(cid:16)
1 − 20γ − 8γ2 + (1 + 8γ)3/2(cid:17)
β2 =
1
8
which is the boundary C2 (25).
,
(B.15)
Article number, page 17 of 23
A&A proofs: manuscript no. secularspin
obtain
ε2 H2 = −
3
4
+ 2X
− 2X
αX 2
M(cid:88)
M(cid:88)
j=1
j=1
N(cid:88)
j=1
E2
j
νjS2
j
νj + αX
ν2
j S2
j
M(cid:88)
ν2
j S2
j
(νj + αX)2
j=1
EjEk cos(θj − θk)
− α(1 − X 2)
N(cid:88)
(cid:104)
M(cid:88)
αX 2
−
3
2
j<k
SjSk
+
Appendix B.4: Third boundary (B/C)
Let us now study the level curve of the Hamiltonian passing
in Σ = −1, which has value −1/2− γ. The procedure is the
same as for the second boundary, and the new formulas
are obtained simply by replacing γ by −γ. There is though
an ambiguity because there are two positive solutions β2
(as a function of γ) which cancel the determinant. The one
corresponding to the bifurcation is the largest, that is,
(cid:16)
1 + 20γ − 8γ2 + (1 − 8γ)3/2(cid:17)
β2 =
1
8
,
(B.16)
which is the boundary C3 (25).
(cid:112)
Appendix B.5: Separatrices
The position at σ = 0 or π of the separatrix emerging from
the hyperbolic point (Σ, σ) = (Σc, π) defines the boundaries
of the resonant region (see Fig. 1). Writing f = F(Σc, π),
the equations to solve are
1
2
Σ2 + γΣ ± β
1 − Σ2 = f .
−
The resolution of this equation requires to square left and
right-hand terms, loosing the information4 about the sign
of cos σ. We obtain a quartic equation in Σ,
(B.17)
1
4
Σ4 − γΣ3 + (γ2 + β2 + f )Σ2 − 2f γΣ + f 2 − β2 = 0 , (B.18)
in which Σc is a double root. It can thus be divided by
(Σ − Σc)2, leading to the quadratic equation
(cid:18) 1
(cid:19)
Σc − γ
Σ
2
Σ2 +
1
4
(cid:18) 3
P2(Σ) =
+
Σ2
c + f + β2 − 2γΣc + γ2
4
(cid:19)
= 0 .
(B.19)
This equation has always two real solutions, provided that
Σc exists (that is, in zones A, B or C). These solutions are
(cid:113)
−β2 + β(cid:112)
Σ± = 2γ − Σc ± 2
where we replaced f by its expression (18) in terms of Σc.
1 − Σ2
c ,
(B.20)
Appendix C: Second-order resonances
Using the intermediary Hamiltonian X = εX1 (31), the
Hamiltonian in the new coordinates is obtained term by
term from Eq. (28). The two first terms are simple: we
have H0 = H0 (given at Eq. 11) and H1 = 0 by definition
of X . The second-order term is more complex since it re-
quires to compute Poisson's brackets. Using of the fact that
{X1,H0} = −H1 and reorganising the terms adequately, we
4 After having computed one solution Σ0, this information is
retrieved by checking the sign of −Σ2
0/2 + γΣ0 − f.
Article number, page 18 of 23
2X(νj + νk) −
j<k
−
α(1 − X 2)νjνk
M(cid:88)
(νj + αX)2 −
νjνkSjSk
+ α(1 − X 2)
(cid:105)
2Xνjνk
νj + αX −
α(1 − X 2)νjνk
(cid:104)
(νk + αX)2
(cid:105)
(νj + αX)2
1
1
2Xνjνk
νk + αX
cos(φj − φk)
cos(φj + φk + 2ψ)
+
(νk + αX)2
ν2
j S2
j
+ α(1 − X 2)
(νj + αX)2 cos(2φj + 2ψ) .
j<k
M(cid:88)
j=1
(C.1)
Since by hypothesis there is no first-order resonance in the
system, the only possible resonant angles at second order
are of the form σ = φj + φk + 2ψ. Let us perform the
canonical change of coordinates
(cid:32) σ
and(cid:32) Σ
γ1
γ2
(cid:33)
(cid:33)
=
=
Γ1
Γ2
−2
1
0
(cid:32)
(cid:32) 1 −1
1
1
0
−1 −2
−1
1
0
1
(cid:33)(cid:32)
(cid:33)
(cid:33)(cid:32) X
−ψ
φj
φk
,
(cid:33)
0
0
1 −3
.
Φj
Φk
(C.2)
(C.3)
Assuming that σ is the only resonant angle, the dynamics
at second order is given by averaging H over all other angles
(this is another change of coordinates close to identity). The
momenta Γ1 and Γ2 become arbitrary constants of motion
that we will conveniently choose equal to zero. Dropping the
unnecessary constants, the resonant Hamiltonian is thus
M(cid:88)
i=1
ν2
i S2
i
νi + αX
νiS2
i − 2X
νj + νk
X
2
M(cid:88)
F(Σ, σ) = −
3
4
−
αX 2
N(cid:88)
i=1
α
2
E2
X 2 −
M(cid:88)
i + 2X
− α(1 − X 2)
(νi + αX)2
i=1
i=1
ν2
i S2
i
(cid:18)
+ α(1 − X 2)νjνkSjSk
1
(νj + αX)2 +
1
(νk + αX)2
(cid:19)
cos σ
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
(C.4)
in which X must be replaced by −2Σ. High-order reso-
nances are quite thin, so it is enough to consider the dy-
namics in the neighbourhood of the resonance centre at first
order:
(average plus oscillating part) for which the expression is
given by (C.1). The homological equation for order 2 is then
{X2,H0} + A2 = A2 ,
(D.4)
which defines the Hamiltonian X2. This leads to
N(cid:88)
(cid:104)
j<k
2X(νj + νk) −
2Xνjνk
νj + αX −
2Xνjνk
νk + αX
3
2
ε2X2 = −
M(cid:88)
+
j<k
−
αX 2
EjEk
µj − µk
sin(θj − θk)
SjSk
νj − νk
α(1 − X 2)νjνk
M(cid:88)
(νj + αX)2 −
+ α(1 − X 2)
j<k
+
M(cid:88)
j=1
+
1
2
α(1 − X 2)
νjνkSjSk
(cid:104)
α(1 − X 2)νjνk
(νk + αX)2
(cid:105)
1
(νk + αX)2
ν2
j S2
j
(νj + αX)3 sin(2φj + 2ψ) .
(cid:105)
1
sin(φj − φk)
sin(φj + φk + 2ψ)
νj + νk + 2αX
(νj + αX)2
(D.5)
We must now compute the remainders at order 3. First of
all, we can simplify their expressions by taking into account
that, by definition: {X1,H0} = −H1, {X2,H0} = −(cid:102)A2 and
{X1,H1} = 2(A2 − H2). We have then
(C.5)
σ =
∂F
∂Σ
= −4αΣ + νj + νk + O(ε2) = 0
νj + νk
,
4α
⇐⇒ Σ0 =
or equivalently X0 = −2Σ0 = −(νj +νk)/(2α). Considering
that X − X0 = O(ε), we have then
(X − X0)2 + αK cos σ ,
F(Σ, σ) = −
in which we dropped the unnecessary constants, and where
(cid:19)
(cid:18)
(C.6)
α
2
νjνkSjSk .
(C.7)
K =
8
(νj − νk)2
(νj + νk)2
1 −
4α2
By injecting the momentum Σ instead of X and by using
the modified time dτ = −4αdt, we obtain
F(Σ, σ) =
This is the Hamiltonian of a pendulum of centre Σ0 and
K. In terms of the obliquity cosine X, the
half width (cid:112)
(Σ − Σ0)2 −
resonance has position X0 and half width 2
(cid:112)
(C.8)
cos σ .
K
4
1
2
K.
Appendix D: Third-order resonances
If there is no resonance at first and second orders, we can
use a canonical change of coordinates close to identity in
order to suppress the angular dependency at first and sec-
ond orders. Let us consider an intermediary Hamiltonian
X = εX1 + ε2X2, such that the new coordinates are given
by its flow at time 1. The Hamiltonian in the new coordi-
nates is then
H = H0 + ε H1 + ε2 H2 + ε3 H3 + O(ε4) ,
where
H0 = H0
H1 = H1 + {X1,H0}
H2 = H2 + {X2,H0} + {X1,H1} +
H3 = H3 + {X1,H2} + {X2,H1} +
1
2{X1,{X1,H0}}
1
2{X1,{X2,H0}}
(D.1)
+
+
1
2{X2,{X1,H0}} +
1
6{X1,{X1,{X1,H0}}} .
1
2{X1,{X1,H1}}
(D.2)
The first-order part of X required to suppress the angular
dependency at order 1 can be directly taken from Eq. (31).
Let us write
A2 = H2 + {X1,H1} +
2{X1,{X1,H0}} = A2 +(cid:102)A2
1
H3 = H3 +
6{X1,(cid:102)A2} +
1
(D.3)
+
2
3{X1,A2} ,
1
3{X1,H2} +
1
2{X2,H1}
(D.6)
Article number, page 19 of 23
A&A proofs: manuscript no. secularspin
ν3
j S3
j
(νj + αX)4 cos(3φj + 3ψ)
ν2
j νkS2
j Sk
Ajk
+ djk + ejk
and:
Ajk =
Bjk =
3(νj + αX)3(νk + αX)3 − cjk
(cid:32)
(νj − νk)2
1
3(νj + αX)2
1 +
αX
νj + αX
+
5
4
α2(1 − X 2)
(νj + αX)2
Cijk = 2νi − νj
(νi + νk)(νi + νj − νk + αX)
(νi − νk)(νj + αX)
(νi + νj)(νi + νk − νj + αX)
(cid:16)
− νk
+ νiνjνk
− eij − eik +
(νi − νj)(νk + αX)
(cid:17)
dji + dki + 2αXbjk + α2(1 − X 2)cjk
.
2 − 2αX xjk − α2(1 − X 2) yjk
3(νj + αX)(νk + αX)
(D.9)
(cid:33)
(D.10)
The expression of the coefficients Dj is very complex. We
will not give it here since they have no interest at this stage
(the angles φj + ψ are non resonant by hypothesis).
As shown in Appendix C, in the pendulum approxima-
tion, the half-width of any possible resonance is two times
the square root of its coefficient divided by α, and its po-
sition is given by the combination of the unperturbed fre-
quencies. Accordingly, the possible resonances at order 3
are gathered in Table 1.
Appendix E: Geršgorin circles
In order to prove that the Lagrange-Laplace matrix for the
orbital inclinations has only negative or zero eigenvalues,
one can use the Geršgorin circle theorem (see Geršgorin
1931 or Varga 2004). This theorem is recalled below, and
we show how it applies to our matrix.
Definition. Let B be a complex N ×N matrix with elements
(bij). The ith "Geršgorin disc" Gi (i = 1, 2..N ) is the closed
disc of the complex plane centred at bii and with radius
bij .
(E.1)
Theorem (Geršgorin 1931). Any eigenvalue of B lies inside
at least one of the Gi discs, i = 1, 2..N.
Corollary. All the eigenvalues of B are located inside the
union of the Gi discs, i = 1, 2..N.
In our case, the matrix B is real (see Eq. 35). It has only
real eigenvalues and one of them is identically equal to zero.
Moreover, given the very particular form of this matrix, the
centre of each Geršgorin disc is located on the real line, with
an abscissa equal to the opposite of its radius. Therefore,
all the eigenvalues of B are negative or zero, as illustrated
in Fig. E.1.
(D.8)
Ri =
N(cid:88)
j=1
j(cid:54)=i
which gives
(cid:105)
M(cid:88)
(cid:104)
j=1
ε H3 =
Dj
cos(φj + ψ)
M(cid:88)
α2(1 − X 2)3/2
M(cid:88)
M(cid:88)
j=1
3
4
−
+ α2(1 − X 2)3/2
j=1
k=1
k(cid:54)=j
(cid:104)
(cid:105)
(cid:105)
(cid:105)
(cid:112)
(cid:112)
+
+
4(νj + αX)4
cos(2φj + φk + 3ψ)
3
(cid:104)
M(cid:88)
−
M(cid:88)
j=1
k=1
k(cid:54)=j
1 − X 2
S2
j Sk
ν2
j νkBjk + νk
+ νj
M(cid:88)
(νj + νk)(νk + αX)
(νj − νk)(νj + αX)
Cijk
M(cid:88)
SiSjSk
(cid:104)
1 − X 2
cos(2φj − φk + ψ)
i=1
j<k
j,k(cid:54)=i
M(cid:88)
× cos(−φi + φj + φk + ψ)
νiνjνkSiSjSk
Aij + Ajk
+ Aik
cos(φi + φj + φk + 3ψ)
(cid:104)
(cid:104)
νiSiEjEk
1
νi + αX
+ α2(1 − X 2)3/2
(cid:112)
+
3
2
αX
1 − X 2
i<j<k
M(cid:88)
N(cid:88)
i=1
j=1
(cid:105)
N(cid:88)
k=1
k(cid:54)=j
1
(cid:105)
µj − µk
cos(φi + θj − θk + ψ)
(D.7)
−
1
where
xjk =
yjk =
zjk =
bjk =
cjk =
djk =
1
νj + αX
+
1
νk + αX
1
(νj + αX)2 +
(νk + αX)2
1
(νj + αX)3 +
1
(νk + αX)3
yjk
νj + νk + 2αX
bjk + zjk
νj + νk + 2αX
1
(cid:32)
3(νj + αX)(νk + αX)
α2(1 − X 2)
(νk + αX)2 +
α2(1 − X 2)
(νj + αX)2 −
+
1 +
2αX
νj + αX −
2αX
νk + αX
(cid:33)
α2(1 − X 2)
(νj + αX)(νk + αX)
ejk = −xjk + 2αX yjk + α2(1 − X 2) zjk
νj − νk
Article number, page 20 of 23
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
Table F.1. Quasi-periodic representation of the orbital dynam-
ics of Mercury.
Fig. E.1. Geršgorin discs in the complex plane corresponding
to the Lagrange-Laplace matrix B for three planets. The centre
of the circles are the diagonal entries of B (red spots). Every
eigenvalue of B lies on the real line, inside the union of all the
discs.
Appendix F: Orbital solution used for the inner
Solar System
In order to apply our method to a given planet, we first
need a quasi-periodic approximation of its long-term orbital
dynamics.
In the case of the Solar System, the search for such series
has been a challenge for centuries, eventually leading to
very complete solutions (up to the degree of chaos inherent
to the system). In the present work, we use the solution
of Laskar (1990), obtained by multiplying the normalised
proper modes z•
i (Tables VI and VII of Laskar 1990)
by the matrix S corresponding to the linear part of the
solution (Table V of Laskar 1990). In the series obtained,
the terms with the same combination of frequencies are then
merged together, finally resulting in 56 terms in eccentricity
and 60 terms in inclination.
i and ζ•
These series are given in Tables F.1-F.4 for the inner
planets, under the form:
z = e exp(i) =
ζ = sin
I
2
exp(iΩ) =
Ej exp(cid:2)i(µjt + θ(0)
j )(cid:3)
Sj exp(cid:2)i(νjt + φ(0)
j )(cid:3) ,
N(cid:88)
M(cid:88)
j=1
j=1
(F.1)
with N = 56 and M = 60. They are used in Fig. 3 of the
present work.
µj ((cid:48)(cid:48)/yr)
5.59644
5.47449
5.71670
4.24882
5.35823
7.45592
4.36906
5.99227
5.65485
6.93423
5.23841
7.05595
7.34103
17.91550
7.57299
17.36469
6.82468
16.81285
3.08952
18.46794
7.20563
17.08266
17.63081
7.71663
28.22069
17.81084
19.01870
17.15752
18.18553
17.72293
18.01611
16.52731
17.47683
16.26122
17.55234
5.40817
18.08627
52.19257
−19.72306
4.89647
0.66708
1.93168
3.60029
−56.90922
53.35188
29.37998
2.97706
−20.88236
28.86795
27.57346
1.82121
27.06140
76.16447
0.77840
51.03334
−0.49216
z
Ej×108
18337396
6902428
5240271
3635276
2815900
2786428
1312738
1035633
998897
934569
829067
634974
235292
165568
164186
157893
77097
70120
60554
48514
48115
43934
39761
29983
21356
15980
15738
15119
14604
12073
10275
9115
7950
7118
6398
6086
5933
3589
2363
922
717
618
447
400
285
169
158
75
70
62
50
45
16
12
9
4
νj ((cid:48)(cid:48)/yr)
(o)
θ(0)
j
−5.61755
110.35
−7.07963
275.01
−7.19493
120.52
−6.96094
30.67
−5.50098
94.89
20.24
0.00000
−6.84091
220.84
−7.33264
113.56
−5.85017
39.22
−5.21610
166.16
−5.37178
272.97
−5.10025
357.62
−6.73842
27.85
−7.40536
335.25
−7.48780
191.47
−6.56016
303.95
−5.96899
14.53
−8.42342
91.98
−3.00557
121.36
9.97 −18.85115
−6.15490
323.91
359.38 −17.74818
−0.69189
202.03
18.14984
273.52
307.83 −18.30007
58.56 −19.40256
39.75 −19.13075
145.02 −26.33023
57.28 −18.01114
48.46 −17.66094
44.83 −17.83857
311.91 −17.54636
80.26 −18.97001
−2.35835
58.89
17.65 −17.94404
120.60 −18.59563
−1.84625
356.17
−4.16482
225.59
113.24 −18.69743
292.23 −18.77933
73.98 −18.22681
39.55 −19.06544
121.40 −17.19656
−3.11725
44.11
−0.58033
134.98
−1.19906
37.61
11.50319
306.81
203.93 −26.97744
212.64 −50.30212
0.46547
223.74
10.34389
146.09
20.96631
38.56
0.57829
323.03
65.10
82.77163
9.18847
136.30
58.80017
164.74
34.82788
−27.48935
−25.17116
−28.13656
ζ
Sj×108
3995819
3015900
1505361
1429554
1424811
1372386
1183049
872607
481844
360659
358805
351141
285961
264351
245583
230801
205822
192248
159813
156874
149031
119892
70222
53922
47541
33322
14506
13964
9092
8620
7011
6248
6027
4687
4398
4035
3435
3303
3167
3104
3100
2777
1298
1067
683
372
341
244
202
196
179
132
68
61
39
36
28
18
17
11
(o)
φ(0)
j
348.70
273.77
105.16
97.95
342.89
107.59
107.89
196.75
165.47
18.91
35.48
195.38
44.50
233.35
47.95
303.47
350.64
211.21
140.33
240.43
89.77
303.28
23.96
111.19
269.86
29.01
125.90
127.29
62.09
318.93
109.13
66.71
253.36
44.73
32.26
278.11
41.72
51.62
41.70
42.83
226.30
230.21
127.26
326.97
17.33
133.87
281.02
44.61
29.83
286.88
191.52
57.78
103.72
128.95
1.15
212.90
294.12
218.53
215.94
314.08
Notes. This representation is used in Fig. 3. It has been directly
obtained from Laskar (1990), see text.
Article number, page 21 of 23
00imaginarypartrealpartTable F.2. Quasi-periodic representation of the orbital dynam-
ics of Venus.
Table F.3. Quasi-periodic representation of the orbital dynam-
ics of the Earth.
A&A proofs: manuscript no. secularspin
z
µj ((cid:48)(cid:48)/yr) Ej×108
2085594
7.45592
1963621
4.24882
1346128
17.91550
17.36469
1164633
659312
5.59644
324058
17.08266
248173
5.47449
239435
16.81285
6.93423
216696
191660
17.63081
188411
5.71670
178786
7.05595
176112
7.34103
129713
17.81084
7.57299
122891
118540
18.18553
116085
19.01870
111520
17.15752
110955
18.46794
5.35823
101244
97999
17.72293
83405
18.01611
67232
16.52731
58638
17.47683
57777
16.26122
6.82468
57706
54138
3.08952
48156
18.08627
47190
17.55234
37236
5.99227
36013
7.20563
5.65485
35915
29809
5.23841
22441
7.71663
20218
4.36906
28.22069
16949
5.40817
3036
−19.72306
1161
1088
0.66708
536
27.06140
470
4.89647
416
29.37998
52.19257
339
277
28.86795
27.57346
244
242
3.60029
−56.90922
216
2.97706
141
69
1.93168
−20.88236
67
63
76.16447
46
1.82121
35
51.03334
0.77840
19
16
53.35188
−0.49216
5
νj ((cid:48)(cid:48)/yr)
(o)
θ(0)
j
0.00000
200.24
30.67 −18.85115
−5.61755
335.25
123.95 −17.74818
−7.07963
110.35
179.38 −18.30007
−5.50098
275.01
−6.84091
274.43
−7.19493
169.77
−6.96094
193.67
120.52 −19.40256
−7.33264
358.98
−3.00557
207.85
58.56 −19.13075
−5.85017
11.47
−0.69189
57.28
−5.21610
219.75
−5.37178
325.02
−5.10025
6.95
94.89 −18.01114
48.46 −17.66094
−6.73842
44.83
131.91 −18.97001
−7.40536
260.26
−5.96899
58.89
194.53 −17.83857
−7.48780
121.36
−6.56016
356.17
197.65 −17.54636
−8.42342
113.56
−6.15490
143.91
39.22 −17.94404
272.97 −18.59563
93.52 −18.69743
220.79 −18.77933
308.38 −18.22681
120.48 −19.06544
113.24 −17.19656
18.14984
73.98
218.72 −26.33023
−2.35835
291.97
−1.84625
217.51
−4.16482
225.73
−3.11725
32.64
−0.58033
43.74
121.40 −50.30212
44.11
11.50319
−1.19906
306.81
93.94
0.46547
203.93 −26.97744
10.34389
143.03
20.96631
148.00
0.57829
316.30
65.10
9.18847
82.77163
135.62
58.80017
164.74
34.82788
−27.48935
−25.17116
−28.13656
ζ
Sj×108
1377170
953835
671575
575205
404368
298364
239467
208804
201837
191673
188959
116984
99208
88217
80983
65885
60616
60304
59016
43620
41358
38337
36654
35440
34592
33636
32924
30942
29977
25773
25048
21099
19361
19259
18879
18853
16888
11902
9063
5577
2677
2187
2022
663
641
215
212
194
184
128
98
82
64
36
24
14
11
7
7
5
(o)
φ(0)
j
107.59
60.43
348.70
123.28
93.77
89.80
342.89
286.39
285.16
277.95
209.10
16.75
140.33
305.90
165.47
23.96
18.91
35.48
195.38
242.09
138.93
224.50
73.36
53.35
350.64
289.13
227.95
123.47
246.71
31.21
89.77
212.26
98.11
221.70
222.83
46.30
50.21
171.81
111.19
127.29
44.72
40.13
51.60
326.97
17.33
29.83
281.02
133.77
286.88
44.89
191.39
57.78
103.72
1.15
128.95
212.90
294.12
218.53
215.94
314.08
z
µj ((cid:48)(cid:48)/yr) Ej×108
1891285
4.24882
1614222
7.45592
1315949
17.91550
17.36469
938579
420011
5.59644
261159
17.08266
197777
17.63081
168931
28.22069
16.81285
168064
161978
6.93423
158097
5.47449
136309
7.34103
134274
7.05595
131495
18.46794
17.81084
126776
120026
5.71670
115855
18.18553
95780
17.72293
95116
7.57299
19.01870
93553
89874
17.15752
81516
18.01611
64497
5.35823
56656
3.08952
56468
16.26122
16.52731
54183
47256
17.47683
47065
18.08627
44664
6.82468
38030
17.55234
27874
7.20563
5.99227
23721
22879
5.65485
18989
5.23841
17369
7.71663
9354
4.36906
52.19257
7041
2871
5.40817
29.37998
1761
1669
27.06140
−19.72306
1591
1259
0.66708
28.86795
1027
902
27.57346
584
53.35188
447
4.89647
76.16447
233
3.60029
233
−56.90922
208
148
2.97706
51.03334
129
70
1.93168
−20.88236
70
1.82121
49
22
0.77840
−0.49216
6
νj ((cid:48)(cid:48)/yr)
(o)
θ(0)
j
0.00000
30.67
200.24 −18.85115
−5.61755
155.25
303.95 −17.74818
−7.07963
110.35
359.38 −18.30007
−5.50098
14.78
−6.84091
128.09
−7.19493
95.11
−6.96094
169.87
275.01 −19.40256
207.85 −26.33023
−7.33264
359.01
−3.00557
187.69
238.56 −19.13075
−0.69189
120.52
−5.85017
237.28
−5.21610
228.46
−5.37178
11.47
−5.10025
39.75
145.02 −18.97001
−6.73842
224.83
94.89 −18.01114
−7.40536
121.36
238.89 −17.66094
−7.48780
311.91
−6.56016
80.26
176.17 −17.19656
−5.96899
194.53
17.65 −17.83857
−8.42342
143.91
113.56 −17.54636
−6.15490
39.22
272.97 −18.69743
93.52 −18.77933
220.76 −18.22681
225.56 −19.06544
120.45 −17.94404
37.54 −18.59563
18.14984
38.70
−2.35835
113.24
−1.84625
73.98
−4.16482
212.64
223.74 −26.97744
−0.58033
134.92
−3.11725
291.91
82.77163
323.03
121.40
58.80017
34.82788
44.11
306.81
11.50319
0.46547
136.30
148.98 −27.48935
−1.19906
203.93
148.46 −25.17116
65.10 −28.13656
10.34389
164.74
20.96631
0.57829
9.18847
−50.30212
ζ
Sj×108
1377263
875509
496020
401987
343071
281401
176869
174079
171242
162618
162229
133519
99249
89258
80968
64554
59814
44770
44540
43589
33642
32525
30484
30067
28903
27932
26251
25667
25550
23507
21866
20949
18500
17676
17327
17304
15500
14745
13530
6694
2098
1981
1812
1074
628
596
581
341
269
191
181
173
167
163
108
85
74
62
36
25
(o)
φ(0)
j
107.59
240.43
348.70
303.28
93.77
269.74
342.89
286.47
285.16
277.95
29.19
127.29
16.75
140.33
125.90
23.96
165.47
18.91
35.48
195.38
253.36
224.50
62.09
53.35
318.93
227.95
123.47
341.65
350.64
109.13
31.21
66.71
89.77
41.70
42.83
226.30
230.21
32.26
278.11
111.19
44.69
39.73
51.59
43.23
17.33
326.97
128.95
212.90
294.12
281.02
286.88
218.53
133.74
215.94
314.08
191.35
57.78
103.72
1.15
29.78
Notes. This representation is used in Fig. 3. It has been directly
obtained from Laskar (1990), see text.
Notes. This representation is used in Fig. 3. It has been directly
obtained from Laskar (1990), see text.
Article number, page 22 of 23
M. Saillenfest et al.: Secular spin-axis dynamics of exoplanets
Table F.4. Quasi-periodic representation of the orbital dynam-
ics of Mars.
z
µj ((cid:48)(cid:48)/yr) Ej×108
4902750
17.91550
4004873
17.36469
2030021
4.24882
16.81285
1853846
1357476
18.46794
1139332
17.63081
1114353
17.08266
706337
28.22069
17.81084
472390
431698
18.18553
399188
19.01870
383488
17.15752
356895
17.72293
303744
18.01611
7.45592
295700
231195
16.52731
210411
16.26122
201641
17.47683
175373
18.08627
17.55234
162274
73627
3.08952
26717
52.19257
25209
6.93423
24970
7.34103
21406
7.05595
7.57299
17424
8182
6.82468
8059
29.37998
8027
27.06140
6750
5.59644
5106
7.20563
28.86795
4795
4212
27.57346
7.71663
3182
3180
4.36906
−19.72306
3055
5.40817
2931
2541
5.47449
2250
53.35188
2008
0.66708
1929
5.71670
1090
76.16447
5.35823
1037
605
51.03334
464
4.89647
381
5.99227
368
5.65485
5.23841
305
3.60029
250
227
1.93168
−56.90922
223
192
2.97706
−20.88236
91
1.82121
65
35
0.77840
−0.49216
10
(o)
νj ((cid:48)(cid:48)/yr)
θ(0)
j
335.25 −17.74818
303.95 −18.85115
0.00000
30.67
91.90 −18.30007
9.91 −17.19656
201.65 −26.33023
359.38 −18.01114
128.11 −17.66094
58.56 −17.83857
57.28 −17.54636
39.75 −19.13075
145.02 −17.94404
48.46 −18.59563
−5.61755
44.83
200.24 −19.40256
−7.07963
311.91
−3.00557
58.89
−0.69189
80.26
356.17 −18.97001
−6.84091
17.65
−7.19493
121.36
−5.50098
225.55
−6.96094
170.37
207.85 −18.69743
359.16 −18.77933
11.47 −18.22681
194.53 −19.06544
−7.33264
37.54
−5.85017
38.70
−5.21610
110.35
−5.37178
143.91
−5.10025
212.64
−6.73842
223.74
−7.40536
93.52
−7.48780
40.88
−6.56016
113.24
−5.96899
120.37
−8.42342
275.01
−6.15490
134.91
73.98 −26.97744
120.52
82.77163
−1.84625
323.03
18.14984
94.89
−4.16482
136.30
58.80017
291.74
113.56
34.82788
−2.35835
39.22
−0.58033
272.97
121.40 −27.48935
192.09 −25.17116
44.11 −50.30212
−3.11725
306.81
203.93 −28.13656
0.46547
149.54
11.50319
65.10
−1.19906
164.74
0.57829
20.96631
10.34389
9.18847
ζ
Sj×108
3464962
1541097
1375324
745752
543058
457927
262761
249135
202620
180575
142530
127098
116625
105161
85957
77142
63897
60870
59221
38863
38505
37498
36566
31116
30502
30460
27285
22307
12681
9492
9443
9241
7310
6758
6278
5900
5417
4915
3922
3471
1993
1457
1419
1278
1169
924
628
592
592
559
458
427
371
170
136
96
59
53
52
33
(o)
φ(0)
j
303.28
60.43
107.59
89.07
154.89
127.29
62.09
318.93
109.13
66.71
305.90
32.26
278.11
348.70
23.33
93.77
140.33
23.96
73.36
286.50
285.16
342.89
277.95
221.70
222.83
46.30
50.21
16.75
165.47
18.91
35.48
195.38
224.50
53.35
227.95
123.47
350.64
31.21
89.77
43.07
128.95
38.22
111.19
51.57
212.90
294.12
44.46
17.33
218.53
215.94
209.84
326.97
314.08
286.88
281.02
133.57
103.72
57.78
191.16
1.15
Notes. This representation is used in Fig. 3. It has been directly
obtained from Laskar (1990), see text.
Article number, page 23 of 23
|
1811.09276 | 1 | 1811 | 2018-11-22T19:00:01 | Morphology of Hydrodynamic Winds: A Study of Planetary Winds in Stellar Environments | [
"astro-ph.EP"
] | Bathed in intense ionizing radiation, close-in gaseous planets undergo hydrodynamic atmospheric escape, which ejects the upper extent of their atmospheres into the interplanetary medium. Ultraviolet detections of escaping gas around transiting planets corroborate such a framework. Exposed to the stellar environment, the outflow is shaped by its interaction with the stellar wind and by the planet's orbit. We model these effects using Athena to perform 3-D radiative-hydrodynamic simulations of tidally-locked hydrogen atmospheres receiving large amounts of ionizing extreme-ultraviolet flux in various stellar environments for the low-magnetic-field case. Through a step-by-step exploration of orbital and stellar wind effects on the planetary outflow, we find three structurally distinct stellar wind regimes: weak, intermediate, and strong. We perform synthetic Lyman-$\alpha$ observations and find unique observational signatures for each regime. A weak stellar wind$\textrm{---}$which cannot confine the planetary outflow, leading to a torus of material around the star$\textrm{---}$has a pre-transit, red-shifted dayside arm and a slightly redward-skewed spectrum during transit. The intermediate regime truncates the dayside outflow at large distances from the planet and causes periodic disruptions of the outflow, producing observational signatures that mimic a double transit. The first of these dips is blue-shifted and precedes the optical transit. Finally, strong stellar winds completely confine the outflow into a cometary tail and accelerate the outflow outwards, producing large blue-shifted signals post-transit. Across all three regimes, large signals occur far outside of transit, offering motivation to continue ultraviolet observations outside of direct transit. | astro-ph.EP | astro-ph |
MORPHOLOGY OF HYDRODYNAMIC WINDS:
A STUDY OF PLANETARY WINDS IN STELLAR ENVIRONMENTS1
John R. McCann1,2 , Ruth A. Murray-Clay2 , Kaitlin Kratter3 , and Mark R. Krumholz4
1Department of Physics, University of California, Santa Barbara, USA; [email protected]
2Department of Astronomy and Astrophysics, University of California, Santa Cruz, USA
3Department of Astronomy and Steward Observatory, University of Arizona, Tucson, USA
4Research School of Astronomy and Astrophysics, Australian National University, Canberra, Australia
(Received September 11, 2018; Revised November 26, 2018)
Submitted to The Astrophysical Journal
ABSTRACT
Bathed in intense ionizing radiation, close-in gaseous planets undergo hydrodynamic atmospheric
escape, which ejects the upper extent of their atmospheres into the interplanetary medium. Ultraviolet
detections of escaping gas around transiting planets corroborate such a framework. Exposed to the
stellar environment, the outflow is shaped by its interaction with the stellar wind and by the planet's
orbit. We model these effects using Athena to perform 3-D radiative-hydrodynamic simulations of
tidally-locked hydrogen atmospheres receiving large amounts of ionizing extreme-ultraviolet flux in
various stellar environments for the low-magnetic-field case. Through a step-by-step exploration of
orbital and stellar wind effects on the planetary outflow, we find three structurally distinct stellar
wind regimes: weak, intermediate, and strong. We perform synthetic Lyman-α observations and
find unique observational signatures for each regime. A weak stellar wind -- which cannot confine the
planetary outflow, leading to a torus of material around the star -- has a pre-transit, red-shifted dayside
arm and a slightly redward-skewed spectrum during transit. The intermediate regime truncates the
dayside outflow at large distances from the planet and causes periodic disruptions of the outflow,
producing observational signatures that mimic a double transit. The first of these dips is blue-shifted
and precedes the optical transit. Finally, strong stellar winds completely confine the outflow into a
cometary tail and accelerate the outflow outwards, producing large blue-shifted signals post-transit.
Across all three regimes, large signals occur far outside of transit, offering motivation to continue
ultraviolet observations outside of direct transit.
Keywords: hydrodynamics, planet -- star interactions, planets and satellites: atmospheres, planets and
satellites: gaseous planets
1.
INTRODUCTION
Atmospheric escape plays a key role in the evolution
of planetary bodies. At their most extreme, processes
that drive escape from the upper atmosphere may sub-
stantially transform the atmospheric composition of a
body over its lifetime. The importance of this mech-
anism is underscored by the recent confirmation that
for short-period exoplanets the radius distribution has a
gap near 1.8 R⊕ (Fulton 2017). The gap's existence was
predicted prior to its discovery as a consequence of the
complete erosion of lower-mass planets' atmospheres by
photoionization-driven atmospheric escape (Owen and
Wu 2013; Lopez and Fortney 2013). While evidence
of atmospheric evolution thus appears imprinted on
planet demographics, observations of contemporary at-
mospheric escape in the regimes governing these popula-
tions are limited, making model validation difficult. Loss
from highly-irradiated planets typically occurs through
hydrodynamic outflows, rather than via the kinetic loss
mechanisms that currently dominate for Solar System
planets. Here, we explore the most approachable systems
for which improved observational constraints on hydro-
dynamic escape can be obtained -- hot Jupiters.
Ultraviolet observations of the hot Jupiter HD 209458
b have found up to a 15 % occultation in the wings of
hydrogen Lyman-α, with effective Doppler shifts of up
1 This pdf contains animated figures (Figures 2, 8, 10, 12, 15,
17, and 19) viewable by doubling clicking in Adobe Acrobat Pro.
Otherwise available at https://gitlab.com/athena ae/athena ae.
to ±150 km s−1, significantly larger than the 5 % occul-
tation observed at optical (Vidal-Madjar et al. 2003;
Ben-Jaffel 2007; Vidal-Madjar et al. 2008; Ehrenreich
et al. 2008). The high occultation and large velocity
are indicative of a fast and extended component of the
atmosphere, interpreted as an escaping planetary wind.
Similar outflows have been reported for another hot and
one warm Jupiter: HD 189733 b (Lecavelier des Etangs
et al. 2010; Bourrier et al. 2013), and 55 Cnc b (Ehren-
reich et al. 20122); and for one hot Neptunian planet, GJ
435 b (Kulow et al. 2014; Ehrenreich et al. 2015; Bour-
rier et al. 2016; Lavie et al. 2017).
Interestingly, the
outflow from GJ 435 b is asymmetric both temporally
and spectrally,3 suggesting a cometary-tail-like outflow
moving rapidly away from the star. Tentative detections
indicate that metals may be present in these escaping
winds, including oxygen (Vidal-Madjar et al. 2004; Ben-
Jaffel and Hosseini 2010), magnesium (Vidal-Madjar et
al. 2013), and carbon and silicon (Linsky et al. 2010;
Lyod et al. 2017). Additionally, hydrogen Hα absorp-
tion has been seen in HD 189733 b's transmission spectra
(Jensen et al. 2012), but its relation to hydrodynamic es-
cape is still uncertain (Barnes et al. 2016). Recently, the
outflow from Wasp-107 b was detected in the 1083 nm
3 Redshifted
2 Along with a non-detection for a super Earth, 55 Cnc e, placing
an upper limit on its mass loss.
(0.7 ± 3.6) % pre-transit
occultation
and (8.0 ± 3.1) % post-transit.
Blueshifted occultation of
(17.6 ± 5.2) % pre-transit and (47.2 ± 4.1) % post-transit (Ehren-
reich et al. 2015).
of
line of excited neutral helium (Spake et al. 2018). This
line, predicted for exoplanet atmospheres by Seager and
Sasselov (2000), and in their outflows by Oklopci´c and
Hirata (2018), provides an opportunity for ground-based
observations.
Current observations have only detected atmospheric
escape for exoplanets with orbital periods less than 20
days. As these close-in planets receive large amounts of
external heating, they are believed to be undergoing hy-
drodynamic escape with outflow structures similar to a
Parker wind (Parker 1958). In an attempt to model the
observations, numerous one-dimensional hydrodynamic
escape simulations have been produced (e.g., Yelle 2004;
Garc`ıa Munoz 2007; Murray-Clay et al. 2009). Yet, out-
flow velocities in these studies generally reach only tens
of kilometers per second, comparable to the sound speed
of the outflowing gas. Note that the temperature of the
gas is limited by radiative cooling to less than a few times
104 K. To resolve inherently asymmetrical processes that
cannot be modeled in 1-D and to investigate interactions
between the outflowing wind and its environment that
may produce high-velocity-offset neutral atoms, multidi-
mensional simulations are needed. A number of authors
have produced such simulations studying pieces of the
problem, including dayside photoionization heating in 2-
D (Owen and Adams 2014) and 3-D (Tripathi et al. 2015;
Debrecht et al. in press), photoionization starting at the
Hill radius (Schneiter et al 2016), stellar wind confine-
ment in 2-D (Stone and Proga 2009; Tremblin and Chi-
ang 2013) and 3-D (Bisikalo et al 2013), magnetic fields in
2-D (Trammell et al. 2014; Owen and Adams 2014) and
3-D (Cohen et al. 2011; Matsakos et al. 2015), and the
possibility of developing a circumstellar torus from the
planetary outflow in global 3-D simulations (Nellenback-
Carroll et al. 2017).
Through these simulations, several possibilities have
been identified that have the potential to enhance occul-
tation in the Lyman-α wings. One option is that given
sufficiently high densities, absorption in the naturally-
broadened line wings may be substantial enough that
a large velocity offset between the gas and the planet
is not required. For example, Trammell et al.
(2014)
demonstrated that HD 209458 b only needed a 50 G
dipole field to explain the observations by producing
a dense and extended equatorial dead zone -- a region
where the outflow's ram pressure is insufficient to over-
come the confining magnetic pressure. Likewise, stellar
wind confinement may increase the column density by
spatially restructuring the outflowing wind, generating
a dense column where significant absorption in the line
wings can occur (Stone and Proga 2009). Alternative
options seek to increase occultation via Doppler broad-
ening by generating a fast neutral population through
additional physics. One such method is the interaction
between a slow neutral planetary wind and a fast ion-
ized stellar wind, which can produce energetic neutral
atoms through charge exchange (Holmstrom et al. 2008).
Furthermore, at the stellar-planetary wind interface, a
Kelvin-Helmholtz instability will lead to stirring that in-
creases the efficiency of charge exchange (Tremblin and
Chiang 2013).
Regardless of whether stellar wind confinement is the
correct or entire explanation for these particular observa-
tions, the stellar environment affects the structure of at-
2
McCann et al.
mospheric outflows, with observational consequences. To
investigate these consequences in 3-D, we take a bottom-
up approach by deconstructing the stellar environment
into three of its individual components -- the ionizing
flux, tidal potential, and stellar wind.
In doing so we
can illuminate how each physical process translates into
observable properties of the outflow, and how one would
expect those signatures to vary for different conditions.
Of the previous simulations, the most complete calcula-
tion of stellar heating in 3-D was Tripathi et al. (2015),
which self-consistently calculated ionization and heating
of the gas. This self-consistent calculation is required to
resolve the ionization structure of the planetary outflow,
which is critical for synthetic observations. Otherwise,
one must rely on previous work that has already done
so for identical parameters, or assume approximate solu-
tions for new parameters.
Yet, while the simulations of Tripathi et al.
(2015)
did include tidal gravity, this study neglected the Corio-
lis force, a stellar wind, and magnetic fields. Expanding
upon Tripathi et al. (2015), we now seek to include stel-
lar winds in a full-rotating frame. A concurrent study
by Debrecht et al. (in press) includes Coriolis force (but
no stellar wind or magnetic fields) and focuses on vary-
ing planet mass and stellar flux. We defer exploration
of magnetic fields, which should play a significant role
in the outflow structure for fields above 1 G (Owen and
Adams 2014), to future work. Though we focus on hot
Jupiters, the concepts explored here should be applica-
ble to a larger demographic, e.g., hot and warm gaseous
planets.
An overview of the goals, model, and setup in this work
are presented in § 2. We provide a detailed description
of the numerical methods and setup in § 3. Results are
given in § 4, followed by a discussion of observational
consequences in § 5. Future work and conclusions are
discussed in § 6.
2. MODEL
We aim to study the interaction between an escap-
ing planetary atmosphere and its host star, in particular
the star's stellar wind, gravitational force, and ionizing
radiation. We will not consider atmospheres undergo-
ing Roche lobe overflow or ablation by the stellar wind.
Rather, our planetary winds will be self-consistently
driven by the energy deposited, via photoionization heat-
ing, in the planet's upper atmosphere. To this end, we
use radiative hydrodynamics to model the evolution of
an atmospheric outflow.
To accurately track the evolution with hydrodynamics,
it is necessary that the outflow remains in the collisional
limit. We confirm this in post-processing with the eval-
uation of the Knudsen number
λ
L
,
Kn =
(1)
where λ = (nσcol)−1 is the mean free path of gas particles
with volumetric number density n, L = (∇ log P )−1 is
the characteristic length scale of the flow, and P is the gas
pressure. The collisional cross section, σcol is obtained
by assuming the responsible collision mechanism for an
individual population: Coulomb scattering for ion-ion in-
teractions, and hardbody collisions for neutral-neutral or
neutral-ion interactions.4
If one is only interested in the planetary evolution, e.g.,
mass-loss rate, it is sufficient for the outflow to remain
collisional only to the sonic surface, where a Mach num-
ber of one is achieved.5 However, we additionally seek
to model the large-scale interaction with the stellar envi-
ronment, which is always governed by the ambient con-
ditions -- modeled here as a stellar wind. Moreover, there
exist ambient conditions which impede the formation of
a sonic surface, in which case the planetary evolution
is also regulated by the ambient conditions. Therefore,
we require the entire outflow to be sufficiently collisional
to be properly modeled by hydrodynamics. Indeed, we
find that within a few sonic radii, both the neutrals and
the ions are well within the collisional limit. Farther
out as the density decreases the neutrals become only
marginally collisional, but are not dynamically or obser-
vationally significant precisely because they reach such
low densities.
In § 2.1 we briefly describe all the physics taken into
account in our models. For §2.2 we discuss the usefulness
of Bernoulli's constant for analyzing these winds. In §2.3
we discuss the physical setup of the problem, and in § 2.4
we give intuition for the relevant length scales in our
flows.
2.1. Physical processes
2.1.1. Radiative transfer for radiative hydrodynamics
Our planetary winds are launched by photoionization
heating from the host star. The frequency (ν) depen-
dent optical depth to ionizing photons along a given ray,
parametrized by s, is given as
τν(s) =
αν(s(cid:48)) ds(cid:48).
(2)
(cid:90) s
s0
Here by definition τν(s0) = 0.
In the work presented
here, the opacities only come from neutral hydrogen ab-
sorption,
i.e., αν(s) = nHI(s) σν,HI, where nHI is the
number density of neutral hydrogen. The near-ionization
frequency-dependent cross section for photon absorption
in neutral hydrogen is approximately (e.g., Draine 2011,
§ 13.1)
(cid:18) ν
(cid:19)−3
ν0
The ionization rate is then given by
I = σHI nHI F0 e−τ .
McCann et al.
(5)
The rate of photoionization heating is the ionization rate
times the energy of the photoelectron released per ion-
ization
G = I Epe.
(6)
Here Epe = hν − IH, where h is Planck's constant and
IH = 13.6 eV is the ionization threshold energy of hydro-
gen.
Within a comoving fluid parcel neutrals are repopu-
lated only via recombination. Since we only follow the
direct stellar ionizing radiation field, and not the diffuse
field generated by recombinations, we will adopt case-B
recombination. This is the appropriate case when opti-
cally thick to ionization, as is the case where the wind is
launched. Farther out in the flow, where the gas is al-
most completely ionized, this may be inappropriate. In
spite of this, we will ubiquitously adopt case-B, making
the recombination rate
R = αB nHII ne.
(7)
Here nHII is the number density of ionized hydrogen, ne
is the number density of electrons, and we use an approx-
imate form for αB = 2.59 × 10−13 (T /104 K)−0.7cm3 s−1
(Osterbrock 1989). Since the atmosphere is solely com-
posed of hydrogen, we take the gas to be electrically neu-
tral so that ne = nHII. We leave the question of metal
cooling and entrainment in the winds for future work.
While our flows near the planet are certainly opti-
cally thick in Lyman-α, it is argued in Appendix C of
Murray-Clay et al.
(2009) that similar flows are suf-
ficiently "thin" that Lyman-α emission is scattered into
the line wings and escapes before being thermalized back
into the fluid via collisions. Therefore, in our cooling
rates we consider both Lyman-α emission from collision-
ally excited neutral hydrogen and radiative recombina-
tion emission
L = Lrr + LLyα.
(8)
σν,HI ≈ 6.3 × 10−18
cm2.
(3)
The rate of energy loss from recombination is given by
(Osterbrock 1989)
For our simulations we implement the Verner et al.
(1996) analytic fits to get more accurate cross sections.
In our study, we will only consider ionizing monochro-
matic light, without any radiation pressure. Therefore,
we will now drop any ν subscripts. Taking the optical
depth from the star to the edge of our simulation to be
negligible, we equate the incoming photon number flux
of the simulation, F0, with the photon number flux of the
host star. Then the flux as a function of optical depth τ
is
F = F0 e−τ .
(4)
4 Neutral-ion interactions are better modeled by induced dipole
scattering and charge exchange, but we ignore those for simplicity
and receive an upper bound for Kn.
5 As, neglecting any magnetic effects, conditions in a sonic region
cannot propagate information to the subsonic region.
3
Lrr ≈ Λrr kB T 0.11 nHII ne.
(9)
The constant Λrr = 6.11 × 10−10 K0.89 cm3 s−1, and kB is
the Stefan-Boltzmann constant. The rate of energy loss
from Lyman-α is given by (Black 1981)
LLyα = ΛLyα nHI ne e−118348 K/T .
(10)
Here ΛLyα = 7.5 × 10−19 erg cm3 s−1. Both the Lyman-
α and recombination cooling are temperature-dependent,
where the temperature is calculated as
T =
P
nkB
.
(11)
n =(cid:80)
The variable n is the total number density of all species,
s ns = nHI + nHII + ne.
2.1.2. Fluid equations
In our simulations we solve the conservative form of
the fluid equations:
+ (cid:126)∇ · (ρ(cid:126)u) = 0,
∂ρ
∂t
(cid:17)
ρ(cid:126)u ⊗ (cid:126)u + P
= −ρ(cid:126)∇φ − 2ρ
(cid:16)(cid:126)Ω × (cid:126)u
(cid:17)
(12)
, (13)
+ (cid:126)∇ ·(cid:16)
∂ρ(cid:126)u
∂t
+ (cid:126)∇ · ((E + P ) (cid:126)u) = −ρ(cid:126)u · (cid:126)∇φ + G − L,
∂E
∂t
∂nHI
∂t
+ (cid:126)∇ · (nHI(cid:126)u) = R − I,
E =
ρ(cid:126)u · (cid:126)u +
1
2
P
γ − 1
.
(14)
(15)
(16)
Here, ρ is the mass density of the gas, (cid:126)u is the bulk
velocity, P is the isotropic pressure tensor with scalar
value P , φ is the mechanical potential, (cid:126)Ω is the frame
rotation vector, and γ is the adiabatic index, for which
we will adopt γ = 5/3. Recall that (cid:126)u ⊗ (cid:126)u is the outer
product of the velocity with itself, sometimes written as
the dyadic product (cid:126)u(cid:126)u.
Equations (12) to (14) have the familiar conservative
forms on the left-hand side, with the relevant source
terms for our problem on the right-hand side. For mass
continuity there are no sources, momentum sources are
external forces arising from the potential and the Cori-
olis force, and for energy there is the change from ad-
vecting through a potential field along with the energy
gained and lost from radiative processes. Recall that the
Coriolis force can do no work. Note that since the cen-
trifugal force can be expressed as the gradient of a scalar
potential, we place it in our mechanical potential, φ, as
discussed in §2.3.1. The continuity equation for neutrals,
Eq. (15), is similar to that for total density with recom-
bination as a source and ionization as a sink. Note that
in Eq. (16) we have used an ideal equation of state for a
perfect gas.
2.2. Bernoulli constant
Bernoulli's constant is a useful tool for analyzing our
simulations, picking a stellar wind, and setting up the ini-
tial conditions of the atmosphere. Lamers and Cassinelli
(1999, § 4.1.1) provide a derivation of the Bernoulli con-
stant for a spherically-symmetric wind; in § A.2 we pro-
vide a generalized derivation along any streamline. For
a reversible ideal gas the Bernoulli constant is
b =
1
2
u2 + h + φ − ∆q,
(17)
where u is the bulk velocity, h is the enthalpy, φ is the
mechanical potential and ∆q is the heat added to the
fluid (see Appendix A). The heat flow along the path is
given by the total local heating rate, G−L, and therefore
∆q can be expressed as
(cid:90)
∆q =
C(s)
G − L
ρu
ds,
(18)
4
McCann et al.
where ρ is the density and C(s) is the streamline, param-
eterized by s.
Something that immediately becomes apparent from
the Bernoulli constant is which parts of the domain are
energetically forbidden. Consider a system in which
there exists a surface defined by φz ≡ b + ∆q. At this
surface, the kinetic energy and enthalpy have gone to
zero (hence the subscript "z" in φz), and all the energy
is necessarily in the potential energy. Thus, the fluid is
bounded by this surface -- the absolute-zero-velocity sur-
face. Therefore, a condition for an unbounded flow is
that (b + ∆q) > φ at all points along its streamline.
We use this criterion to help determine our initial condi-
tions, picking only bounded planetary atmospheres and
unbounded stellar winds.
When a flow is unbounded it is meaningful to talk
about its asymptotic velocity. For frames where both
φ → 0 and h → 0 as r → ∞, the asymptotic velocity can
be calculated as
u∞ =(cid:112)2 (b + ∆q∞).
(19)
For an ideal gas, h → 0 is equivalent to T → 0. In prac-
tice, as long as the flow is supersonic as r → ∞, Eq. (19)
is approximately correct as u2∞ (cid:29) h∞ ≈ c2
s,∞, where cs,∞
is the sound speed as r → ∞. Here, the notation ∆q∞
reminds us that it includes all the energy injected into
the wind out to infinity, where we have assumed that the
differential heat flow is zero.
2.3. Physical setup
2.3.1. Reference frame
We place our planet on a circular orbit and adopt a
rotating reference frame in which the planet and star are
at fixed locations. This frame centered on the barycenter,
has the rotation vector given by the Third law of Kepler
(cid:126)Ω =
G(Mp + M(cid:63))
a3
z.
(20)
Here G is the Newtonian gravitational constant, M(cid:63) is
the primary mass, Mp is the secondary mass, a is the
semi-major axis of the secondary's orbit around the pri-
mary, and the z -- axis is the axis of rotation. In such a
frame the static potential is given by
− 1
2
φ = − GMp
− GM(cid:63)
r(cid:63)
Ω2 r2⊥,
(21)
r
where, for a given point, r is the distance to the sec-
ondary, r(cid:63) is the distance to the primary, and r⊥ is
the distance to the barycenter projected into the orbital
plane.
2.3.2. Initial Planetary atmosphere setup
Within this reference frame we place a point mass to
simulate the planet's core, on top of which sits an at-
mosphere. As the wind is launched in the upper atmo-
sphere of a planet, the gas is relatively dilute and warm
enough to be well described as an ideal gas. We ignore
any viscous dissipation, and take all processes to be re-
versible. Absent any external energy input, the most
stable solution of an atmosphere will be an adiabat. We
thus construct an isentropic atmosphere that satisfies the
(cid:114)
polytropic relationship, P = KρΓ, with an polytropic ex-
ponent, Γ, equal to the heat capacity ratio, γ.
In reality the bolometric flux from the star drives
the upper atmosphere of a planet to be approximately
isothermal at the skin temperature (See § 3.6 of Pier-
rehumbert 2010). Due to a lack of incorporating the
bolometric flux in our simulation, and since Γ ≈ 1 is not
an appropriate approximation throughout our simulation
domain, e.g, in the ionized outflow, we use an isentropic
atmosphere with the temperature at the base of the wind
equal to the planet's skin temperature.
McCann et al.
Our atmosphere is contained well within the planet's
Hill sphere, so for now we can ignore stellar gravity when
calculating its analytical initial profile. Note that in
practice we will use the full potential so that the at-
mosphere is not technically spherically symmetric, but
for simplicity of discussion we will assume such symme-
try.6 Furthermore, we consider planets that are tidally
locked to their host stars, and will ignore any rotational
effects tending to make the atmosphere oblate. Under
these conditions we require the atmosphere to be in hy-
drostatic equilibrium. We ignore the gravitational effect
of the gas itself, both analytically and numerically in our
simulations.7 From Eqs. (17), (A14), and the adiabatic
equation of state, an adiabatic atmosphere that is every-
where static has a density profile
(cid:20)
(cid:21)1/(γ−1)
(φp − φ(r))
hp
ρ(r) = ρp
1 +
.
(22)
Here variables with subscript "p" denote their value at
the planetary surface, Rp, which we define to be the ra-
dius where τ = 1. We emphasis that, τ = 1 at Rp is only
the initial condition at time step zero in our simulations
of our atmosphere. The optical depth to ionizing pho-
tons at Rp is not fixed during the simulation, as the wind
self-consistently picks its base. We pick the atmosphere's
Bernoulli constant such that the atmosphere is bound,
which is found by solving the equation φ = φp + hp.8
Since our atmosphere is (nearly) spherically symmetric
we shall call this the zero radius, Rz, as all variables go to
zero at this surface. This is often called the homentropic
atmosphere since it is isentropic, with constant entropy
along rays, and is spherically symmetric -- thus, of "the
same entropy" everywhere throughout. For an ideal gas
in a point mass potential this is equivalent to
ρ(r) = ρp
1 +
(γ − 1)GMp
γ c2
s,p
r
− 1
Rp
(cid:18) 1
(cid:19)(cid:21)1/(γ−1)
(cid:20)
Here cs,p is the isothermal speed of sound at Rp.9
6 See Figure 1, which shows the solution for the full non-spherical
potential. Yet, by visual inspection the atmosphere is virtual spher-
ically symmetric.
7 Thus while the atmosphere has a polytropic equation of state,
it is not a polytrope as it is not a solution to the Lane-Edmen
equation, i.e. no self-gravity, and there are an infinite number of
analytic solutions.
8 Thus the relevant parameters are the radius, mass and skin
temperature of the planet. For our simulations we pick parameters
of numerically convenient hot Jupiters, see § 3.2.2 for more details.
9 Notice that from this equation, and the limit definition of e,
s,p (1/r − 1/Rp)), which is the
the limγ→1 ρ(r) = ρp exp(GMp/c2
profile of an isothermal atmosphere.
Figure 1. The initial density profile of a hydrostatic atmosphere.
Within the innermost circle, r < Rmask, the density is held to
a constant value, ρmask. Between Rib and Rmask the density is
held fixed at every time step to its original analytic hydrostatic
solution. The radius of the planet, Rp, is where τ = 1 initially.
The outermost circle, Re, is the edge of the atmosphere where the
density tends to zero, past which is a low density, pressure matched
ambient medium.
Numerically fixed inner-boundary conditions, which
are extrapolated from Eq. (22) with the initial condi-
tions at Rp, are set at an inner boundary radius. The
inner boundary, Rib, is deep enough within the atmo-
sphere that conditions there do not affect the outflow.
While we setup an initial atmosphere with predeter-
mined conditions, the stellar flux boundary condition
self-consistently evolves the atmosphere above τ = 1.
Therefore, below τ = 1 our atmosphere is relatively
static, and our model should not necessarily be consid-
ered valid there -- even though it is far outside our inner
boundary. Justification for the insensitivity of the solu-
tion above τ = 1 to what lies below is given in Appendix
A of Murray-Clay et al. (2009).
See Figure 1 for the initial atmosphere with labeled
radii. We note that our choice of Rp as the radius at
which τ = 1 is not the same as the radius of the op-
tical surface of the planet (as probed, for example, in
optical transit measurements). We make this choice for
Rp because our model is only valid beyond this radius.
For typical hot Jupiters, the difference between the radii
of the optical surface and the τ = 1 surface to ionizing
photons is of order 10 % (Murray-Clay et al. 2009).
. (23)
2.3.3. Ambient medium setup
Ideally, conditions outside of the atmosphere would be
those of a typical interplanetary medium, which for most
star systems takes the form of a stellar wind. However,
as we wish to break the stellar environment down into
its components, in some simulations we do not include
a stellar wind. When the stellar wind is absent, we in-
stead use an ambient medium that is as low-density as
numerically feasible and that is pressure matched to the
outer edge of the planetary atmosphere. For numerical
reasons we truncate the planetary atmosphere just prior
to the zero radius (defined in § 2.3.2) at a radius we call
the atmospheric edge, Re.
Initially it is important that the ambient medium is
both pressure matched to the numerical edge of our at-
5
ReRpRibRmaskmosphere, and that the ambient medium is itself pressure
supported so it does not collapse onto the planet. This is
accomplished with another hydrostatic atmosphere sim-
ilar to Eq. (22)
(cid:18)
(cid:19)1/(γ−1)
φa − φ(r)
ha
ρ(r) = ρa
1 +
,
r > Re.
(24)
Here the subscripts "a" denotes an ambient medium ref-
erence variable, where the reference location is the am-
bient medium/planetary atmosphere interface (Re).
By design the ambient medium's reference density ρa <
ρe -- the density of the atmosphere at Re. Thus by also
requiring pressuring matching at the interface (φa = φe),
Pa = Pe, so that ha > he. Since φz,a = φa + ha deter-
mines the ambient absolute-zero-velocity surface, then
the zero radius of the ambient medium is larger than the
planetary atmosphere's bounding surface.10 Due to nu-
merical considerations, a balancing act between too high
or too low of values for ha (or relatedly ρa) occurs and is
discussed in Appendix B.
2.3.4. Stellar wind setup
When we seek to include the interplanetary medium
in the form of a stellar wind, we set a stellar wind in-
flow boundary condition and initialize our domain with
that of a steady-state stellar wind instead of an ambient
medium. One challenging aspect of our simulation is the
injection of a realistic stellar wind. As our energy depo-
sition into the planetary atmosphere is detailed, we use
an adiabatic index of Γ = γ = 5/3 throughout the entire
simulation. This differs from numerous previous simula-
tions only in that we have no "hidden" energy injection
or ad hoc redistribution built into our adiabatic index,
such as an isothermal Parker Wind with Γ ≈ 1 (e.g.,
Stone & Proga 2009; Tremblin & Chaing 2013; Caroll-
Nellenback et al. 2017). This means our fluid behaves
isentropically in the absence of ionizing radiation.
Typically, stellar wind models that do not resolve the
heating sources use such "polytropes" (Γ ≈ 1). Unfor-
tunately for our simulation, as discussed in § A.3, there
is no transonic wind solution for a polytropic fluid with
polytropic index Γ = 5/3. Note that, stellar wind mod-
els which do resolve heating terms also use Γ = γ = 5/3,
but include things such as: heat conduction, Alfv´en wave
dissipation, resistive and viscous dissipation, or coronal
heating sources (Miki´c et al. 1999). As our only en-
ergy injection is from ionization, we cannot generate an
optically thin transonic stellar wind without another en-
ergy injection method, e.g., via magnetic fields as seen
in Alfv´en-driven winds (Lamers & Cassinelli 1999, Ch.
10). Fortunately, by using the Bernoulli constant we can
set the boundary conditions to induce a stellar wind that
within the domain of our simulation mimics a transonic
wind locally. The only catch is that the stellar wind is
always sub- or supersonic (see § 3.2.2 for more details).
Therefore, our stellar wind is modeled as a spherical
isentropic wind,11 i.e., a "polytrope" with index Γ = γ
10 Since φz,a = φa + ha and φz,p = φp + hp = φe + he, but
ha > he and φa = φe, then φz,a > φz,p or Rz,a > Rz,p.
11 Necessarily due to the fact that there is no energy injection
mechanism for the optically-thin stellar wind, and that the entire
work done in our hydrodynamic simulation is reversible.
McCann et al.
(see Footnote 7 and § A.3), for which the velocity pro-
file is derived in § A.3. Note that spherical symmetry
may not be realistic, even at close distances, but we will
make this assumption regardless (Vidotto et al. 2018).
Therefore, we use Eq. (A18) with Γ = γ to evaluate the
stellar wind's velocity structure. Notice that since u(r)
is independent of ρ. Therefore, given a u(cid:63),0 and T(cid:63),0 at
r(cid:63),0 (a reference radius not necessarily equal to the stel-
lar radius), we can adjust the total pressure of the stellar
wind by scaling the stellar proton number density, n(cid:63),0,
while leaving all other profiles unaltered
P(cid:63),total(r) = ρ(cid:63)u2
(cid:63) + P(cid:63)
(cid:32)
(cid:32)
= n(cid:63),0
(mHII + me)
(cid:33)
u(cid:63),0 r2
(cid:63),0
u(r) r2
(cid:32)
u(r)2
(cid:33)γ(cid:33)
+ 2kBT(cid:63),0
u(cid:63),0 r2
(cid:63),0
u(r) r2
.
(25)
Here me is the mass of an electron and mHII is the mass
of an ionized hydrogen. Note that there is a factor of two
on the thermal pressure to account for both the protons
and electrons. Thus, we only need to alter n(cid:63),0 to tune
the stellar wind strength. Controlling n(cid:63),0 grants us a
handle on where the bow-shock interface of the planetary
and stellar winds occurs, and on whether the planetary
outflow is a wind or breeze. Moreover, this allows us
to use the same velocity and temperature profile, and
therefore Mach profile, across all of our stellar winds.
We note that realistic stellar winds do not merely differ
between one another by a density scaling, and that we
are only probing a small area of the possible phase space
of stellar winds.
In this study we focus on the effects
of a confining stellar wind, and therefore only require
a handle on the total wind pressure to make compar-
isons between wind strengths. Studies probing charge
exchange, or stellar magnetic fields, will require more
realistic modeling of stellar winds. For a review of real-
istic stellar-wind numerical modeling see Gombosi et al.
(2018).
2.4. Length scales
To determine an appropriately sized domain for any
simulation,
it is important to understand the length
scales of the problem to ensure that everything can be
captured within the domain. Additionally, length scales
offer insight into the resulting structures and their ori-
gins. For our problem, the important scales for the out-
flow are: the outflow's scale height (the importance of
which has already been discussed in § 2), the optical
depth one surface for ionizing photons, the sonic point
where the planetary outflow transitions from subsonic to
supersonic, the planet's Hill radius, the Coriolis length of
the planetary outflow, and the bow shock radius where
the planetary and stellar winds collide.
Optical depth one to ionizing photons: While τ = 1
along the substellar ray was initially set at Rp, we allow
the simulation to self-consistently choose where τ = 1 as
the flow evolves. This will depend on the incoming ion-
izing flux, recombination rate, and advection of neutrals
in the planetary outflow. However, one can still chose np
6
such that τ = 1 ends up near Rp. This choice depends on
the strength of the ionizing flux, and is briefly discussed
in Appendix C.
Sonic point: The sonic point is where the outflow goes
transonic ((cid:126)u = cs and (cid:126)u · (cid:126)∇(cid:126)u > 0). However, for a
self-consistently launched outflow with ionization heat-
ing its location is difficult to know a priori. For ref-
erence, the sonic point radius for an isothermal Parker
wind is rs,iso = GMp/c2
s . This expression can be used to
approximate the location in a self-consistently launched
outflow by evaluating cs using the temperature is at the
base of the wind, T ∼ 104 K. For a fuller understanding
of how the sonic point depends on ionization heating see
the discussion in § 2.2.3 of Murray-Clay et al. (2009).
Effective Hill radius: The Hill radius, where stellar
tidal gravity balances the planet's gravity, is well ap-
proximated as
(cid:114) Mp
3M(cid:63)
rH = a 3
.
(26)
However, the Hill radius is derived for particles that do
not experience a pressure force. Just as the pressure
force causes gas in a protoplanetary disk to orbit at sub-
Keplerian velocities by effectively reducing the central
mass, here the pressure force reduces the Hill radius by
effectively reducing the planet's mass. We call this radius
where the planet, tidal, and pressure forces balance the
"effective Hill radius," which will be interior to Eq. (26)
for monotonically decreasing pressure profiles.
Coriolis length: Now imagine a ballistic particle only
experiencing the Coriolis force, (cid:126)a = −2(cid:126)Ω × (cid:126)v. The par-
ticles moves in a circle with a period of t = π/Ω, since
the frequency of the acceleration is twice the frame ro-
tation, 2Ω, and a full revolution is 2π. Recall that the
Coriolis force does no work,
serves to transfer momentum between coordinate axes.
Therefore, we arbitrarily define when a particle has been
significantly affected by the Coriolis force as when its ini-
tial momentum has been deflected by one radian, i.e., it
has traversed one radius -- the length scale of the circle.
Thus the time scale on which particles are significantly
affected by the Coriolis force is (2Ω)−1. Given the aver-
age velocity over that trajectory, (cid:104)v(cid:105), the Coriolis length
scale is
(cid:17)· (cid:126)v = 0, and only
(cid:16)−2(cid:126)Ω × (cid:126)v
LΩ =
(cid:104)v(cid:105)
2Ω
.
(27)
One could similarly define the Coriolis length to be the
scale on which the outflow reaches a Rossby number of
one, Ro ≡ v/(2ΩLΩ). For an outflow with its sonic point
interior to the Hill sphere, (cid:104)v(cid:105) is on the order of u∞ from
Eq. (19).
Bow shock : The bow shock radius, Rbow occurs where
there is a pressure match between the stellar wind and
the planetary outflow
(cid:2)P(cid:63) + ρ(cid:63)u2
(cid:63),⊥ = P + ρu2⊥(cid:3)Rbow.
(28)
Here the "(cid:63)" subscripts denote the stellar wind, which
is not constant throughout the domain. Subscript "⊥"
denotes the normal component of the velocity to the bow
shock interface. As it is the stellar wind that shrouds the
7
McCann et al.
planet, the bow shock is roughly spherical with respect
to the planet's origin between the planet and star, but
asymptotes to a radial line far past the planet.12 Since we
can analytically solve for the stellar wind structure, we
can numerically solve for the location of the standoff ra-
dius given the numerical structure of the planetary wind
in the absence of a stellar wind. We note that because
the shocked stellar wind does not radiate effectively, the
width of the shocked stellar wind region can be substan-
tial, and care must be taken to choose a box size large
enough to enclose the shock.
3. NUMERICAL METHODS
3.1. Athena
To solve the model described in § 2 we use the
publicly available magnetohydrodynamics code Athena
v4.2 (Stone et al.
2008). This Eulerian code has
been rigorously tested and highly parallelized, making an
ideal starting point to solve our radiative-hydrodynamic
model. Two additional packages are utilized to add
physics beyond the ideal fluid equations. The first is
the ionization package from Krumholz et al. (2007), up-
dated to incorporate static mesh refinement (SMR) for
plane-parallel ionization (Tripathi et al. 2015). The ra-
diative transfer in this package is operationally split from
the hydrodynamic update and is done by radiative sub-
cycling between hydro time steps. Secondly, Athena's
shearing-box physics package (Stone et al. 2010) is used
to implement the Coriolis force, as described in § 3.1.1.
Standard fluid algorithmic choices are as follows. We
use the piecewise-parabolic method, a third-order spa-
tial reconstruction method, for reconstructing the fluid
variables at cell interfaces. For the Riemann problem of
the interface fluxes, Roe's linearized solver is used. Our
integrator is the 3-D directionally unspilt corner trans-
port upwind (CTU) scheme. Static mesh refinement is
used around the planet to ensure scale heights within the
fluids are well resolved. As discussed in Tripathi et al.
(2015), we use the H -- correction algorithm to avoid the
carbuncle instability from the wind's convergences on the
nightside.
3.1.1. Modifications and use of Athena's features
In order to run the simulations successfully and to
improve their accuracy, we modify the default Athena
code and implement a few non-standard features. First,
we added first-order flux correction to the corner trans-
port upwind (CTU) integrator, using the same method
as already implemented in Athena's van Leer integrator
(Beckwith and Stone 2011). This scheme detects when
calculation of the flux at higher order leads to a negative
density or pressure, and self-consistently redoes the flux
calculation at the boundary of each affected cell using
more diffusive, and hence more stable, first-order fluxes.
Resorting to first-order fluxes turns out to only be nec-
essary initially while the wind as launching, for similar
reasons as discussed below regarding our prolongation
slope limiter changes.
Second, we modified Athena's shearing box physics fea-
ture so that the Coriolis force without a centrifugal term
12 See Figure 10 of Murray-Clay et al. (2009) for a cartoon of
the geometry.
can be used with non-periodic boundary conditions. To
do so without adding additional fictitious forces one can
set the shear parameter to zero, q = −∂log Ω/∂log r = 0.
A shear parameter of zero suggests solid body rotation,
for which central forces are balanced everywhere and the
only force felt is the Coriolis force. Thus the only ficti-
tious force the shearing-box module adds is the Corio-
lis force, while the tidal forces are taken care of in the
static potential without any of the usual shearing-box
approximations. Additional steps need to be taken as
the shearing-box approximation assumes periodic bound-
ary conditions, with Athena's shearing-box feature hard-
coded to remap boundaries without consideration to the
user-input boundaries.
As some simulations are in a rotating frame, issues can
arise when a flow is bending near a boundary. To pre-
vent unphysical inflow from being extrapolated from the
boundary conditions, we use what are sometimes referred
to as dipole boundary conditions. That is to say, we use
the standard outflow boundary conditions when the bulk
velocity normal to the boundary is outwards, but restrict
the mass inflow to the initial ambient medium when the
normal velocity is inwards. While this sufficiently re-
duces undesired inflow, there are still reflections at the
boundary that are not damped and can lead to oscilla-
tions. While these oscillations are present in our sim-
ulations, they occur where the density and pressure are
orders of magnitude smaller than those around the planet
in the domain of interest. Note that these are not stan-
dard Athena boundary conditions, and are implemented
as user defined boundary conditions.
Next, we also found it necessary to add an additional
slope limiter to Athena's prolongation operator as ap-
plied at SMR boundaries. Since radiative cooling in our
simulations is rapid, when the ionization front in the
planetary atmosphere first expands it is characterized by
a very large density jump. When this structure passes
over SMR boundaries, the prolonged density or pressure
can be negative unless it is appropriately limited. We
therefore limit the slope in each direction to be less than
4/3,13 which is sufficient to ensure that the prolonged
quantities remain positive definite. This limit only mod-
ifies the simulation during the early phase of evolution
when the outflow is expanding, and does not apply in the
steady-state configuration reached at later times as this
structure has passed over all of our SMR level bound-
aries.
Finally, we made minor improvements to Krumholz et
al.'s (2007) ionization package by modifying the temper-
ature calculation to produce higher accuracy. Our mod-
ified version of Athena, other tools used in the produc-
tion of this paper, a version of this paper with embedded
movies, and more are freely available for download and
use at our GitLab repository.14
3.2. Domain setup
3.2.1. Reference frame
All simulations have the planet at the origin, and de-
pending on the simulation, we adopt either an inertial or
13 This limit comes from the cell-centered distances between
child cells and their parents, 1/4, and the number of dimensions,
3.
14 https://gitlab.com/athena ae/athena ae
8
McCann et al.
a non-inertial rotating reference frame. By using a frame
centered on the secondary we can use a smaller domain,
which keeps both total computational expense and error
from using plane-parallel radiation to a minimum. For
rotating frames, let the primary be located at (cid:126)r(cid:63) = −a x
and the barycenter at (cid:126)rb = −aM(cid:63)/(M(cid:63) + Mp) x, so that
the potential in Eq. (21) is recast as
φ((cid:126)r) = − GMp
(cid:126)r − GM(cid:63)
(cid:126)r − (cid:126)r(cid:63) − 1
2
Ω2(cid:126)r⊥ − (cid:126)rb2.
(29)
The first term is included in all of our simulations as
the planet is always present, while the second and third
term are only included in simulations where tidal po-
tential is considered. In Athena this is implemented by
setting the static potential function to Eq. (29). This
allows us to examine the tidal forces aspect of a rotating
frame without considering the Coriolis force, which can
be included with the shearing-box package described in
§ 3.1.1.
3.2.2. Fluid initial conditions
Within the simulation is a planetary atmosphere sur-
rounded by an ambient medium. The parameters that
characterize the planet and its atmosphere are the
planet's mass, Mp, and radius, Rp, and the temperature,
Tp, at Rp. While the mass and radius can be treated
as free parameters used to model any desired planet,
to launch a wind from ionization heating the planet's
temperature must be well below that of the launched
wind (Appendix A of Murray-Clay et al. 2009). Thus
the temperature sets a physical scale within the prob-
lem (Tp (cid:28) 104 K for hot Jupiters), making planets with
larger escape speeds or smaller wind temperatures, which
have a smaller scale height to planetary radius ratio,
more challenging to simulate.
Other atmospheric quantities can be calculated from
these primary parameters and certain assumptions. One
such assumption -- accurate for ionizing fluxes studied
here -- is that planetary escape occurs in the energy-
limited regime rather than the recombination-limited
regime, as described in Murray-Clay et al.
(2009).
Therefore, the surface of optical depth unity to ionizing
photons does not move appreciably between the start of
the simulation and when the wind has reached steady
state. As the initial isentropic atmosphere's density and
pressure scale heights are independent of the particle
number density, we can always set the location of the
optical depth unity surface to be at Rp by scaling the
number density. We do so by setting
np =
where (cid:101)H is a scale height defined in Appendix C. Oth-
erwise, in the recombination-limited regime one could
balance recombination with ionization, to get a proper
number density such that detailed balance is achieved
near Rp.
Setting aside the temperature, a planet's mass and ra-
dius will determine the scale height and in turn the simu-
lation resolution required. For an isentropic atmosphere,
the scale heights become infinitesimally small near the
σHI(cid:101)H(Rp)
1
,
(30)
edge of the atmosphere and at the core. This does not
matter near the atmospheric edge as we truncate before
reaching the zero radius, and once the simulation begins
to run the edge will be replaced by a wind in any event.
However, the singularity near the core requires that we
adopt an inner boundary in our simulation to avoid nu-
merical difficulties.
To set the inner boundary, Rib, we require a few scale
heights between τ = 1 and Rib, so that the base of the
wind will be self-consistently found without interference
from the inner boundary. Here we define a scale height
to be one e -- folding in density (see Eq. (C1)), and choose
Rib to be two scale heights below the initial atmosphere's
τ = 1. We then set Rmask five cells below Rib (number
of ghost cells plus one), within which we hold the all
variables to the same fixed value. By resetting all the
cells between Rib and Rmask at every time step to their
hydrostatic solution, we create an internal boundary con-
dition, as no information within Rmask can propagate out
by construction. By using a masking radius instead of a
softening radius (i.e., a Plummer radius) in our potential
we avoiding introducing artificial errors in our potential.
Past the numerical edge of the planetary atmosphere
is the ambient medium. When the ambient medium is
not supposed to represent a stellar wind, our goal is to
minimize its impact on the planetary wind. The de-
tailed structure of our ambient medium is described in
Appendix B, but it suffices to say that by using a low
density, initially pressure-matched, ambient medium, we
prevent infall onto the atmosphere and stop the plane-
tary wind from entering a snowplow phase. When we do
include a stellar wind, we initialize the ambient medium
with the velocity structure given by Eq. (A18). With
the velocity structure and a stellar mass-loss rate we can
then calculate the density and temperature profiles of the
wind with Eqs. (A16) and (A17) respectively. To indef-
initely sustain the wind, the same formulation used to
refresh the planet's hydrostatic lower atmosphere with a
fixed inner boundary is used. That means within a fixed
inner boundary radius, r(cid:63),ib = 4 × 1011 cm, centered on
the star's origin, stellar wind conditions are held constant
throughout time. Lastly, the domain's dipole boundary
conditions (§ 3.1.1) are then modified to respect these
conditions.
As mentioned in §2.3.4, the stellar wind is either always
sonic or sub-sonic due to the stellar wind's polytropic in-
dex of Γ = 5/3. Therefore, we chose a sonic stellar wind,
such that it will shock on the planetary outflow. While
the wind is technically isentropic, it has been carefully
chosen to mimic a more realistic stellar wind within the
domain of the simulation. Specifically, the stellar wind
starts at r(cid:63),0 with v = 200 km s−1, but by the time it
reaches the planet it has accelerated to v = 290 km s−1
(similar stellar outflow velocities are found, e.g., in a stel-
lar wind model for HD 219134 by Vidotto et al. (2018)).
It continues to accelerate past the planet with a maxi-
mum of nearly 300 km s−1. We note that outside of our
simulation box, the isentropic stellar wind is not a good
model for the stellar wind profiles.
3.2.3. Stellar radiation
A planetary wind is launched by irradiating our com-
putational domain with monochromatic plane-parallel
radiation. Ionizing flux enters from the negative x -- axis,
9
Table 1
Simulation parameters
Parameter
Planet
Mass, Mp (g)
Radius, Rp (cm)
Temperature at Rp, Tp (K)
Density at Rp,# ρp (g cm−3)
Orbital Parmeters
Semi-major axis, a (cm)
Orbital period,# P (s)
Star
Mass, M(cid:63) (g)
Radius, R(cid:63) (cm)
Ionizing flux, F0 (cm−2 s−1)
Photon energy, hν (eV)
McCann et al.
Value
5.0×1029
1.5×1010
1.1×103
6.64×10−16
1.0×1012
5.5×105
1.989×1033
6.957×1010
2.0×1013
1.6×101
Stellar wind
4.0×1011
Reference radius, r(cid:63),0 (cm)
1.35×106
Temperature at r(cid:63),0, T(cid:63),0 (K)
2.0×107
Velocity at r(cid:63),0, v(cid:63),0 (cm s−1)
Proton number density at r(cid:63),0, n(cid:63),0 (cm−3) (1.5, 15, 70)×103
(1.3, 13, 59)×10−16
M(cid:63) (M(cid:12) yr−1)
Mass-loss rate,#
(1.3, 13, 61)×10−7
Pressure at planet,# P(cid:63),total(a) (dyn cm−2)
Note. -- Parameters used in simulations. Parameters that are
not free parameters, but are determined by other variables are de-
noted with a # superscript. Tuple values in parentheses correspond
to the weak, intermediate, and strong stellar winds accordingly.
altering the ionization state and depositing energy into
the fluid. To prevent transients from impacting the early
simulation, the flux is ramped up so that the wind is
gently launched into the ambient medium. Let the true
physical stellar flux be equal to our flux at time equal
infinity, F (∞) = F∞. Our initial flux will start at a
factor f0 of the physical flux, such that F (0) = f0F∞.
We gradually increase the flux using a Gaussian rate of
change ramp function with standard deviation σ, so that
after time t the flux is
(cid:18)
(cid:18)
(cid:18) t
t1/2
(cid:19)(cid:19)(cid:19)
.
− 1
F (t) =
F∞
2
1 + erf
erf
−1 (1 − 2f0)
(31)
Here erf() is the error function and erf−1() is the inverse
error function. The halfway point of the flux ramping
−1 (1 − 2f0), such that F (t1/2) =
occurs at t1/2 =
2 F (∞). In practice, once F (t) = 0.999F∞ we set F (t) =
1
F∞ for the rest of the simulation duration.
2σ erf
√
3.3. Parameters used
To efficiently model atmospheric escape around hot
Jupiters, we pick parameters that enable good resolu-
tion at reasonable cost while still being physically moti-
vated. For the planet we use Mp = 5.0 × 1029 g ≈ M(cid:89),
Rp = 1.5 × 1010 cm ≈ 2R(cid:88), and Tp = 1100 K with a
semi-major axis a = 1012 cm ≈ 0.07 au. Recall that (cid:89) is
the astronomical symbol for Saturn and (cid:88) the symbol
for Jupiter. Our choices correspond to a large planetary
radius, which helps increase the scale height by decreas-
ing the local surface gravity g. For a full understanding
of how these parameters affect the scale height, see Ap-
pendix C. Our parameters are close to those of Wasp-17
b, one of the more extreme exoplanets. We run all of our
simulations for 2 × 106 s, which is over 3.5 orbital peri-
ods. After which, all simulations appear to have reached
a steady or quasi-steady state.
The domain varies depending on which simulation we
run. As explained in § 4, we run three classes of simula-
tion depending on which physical processes we include;
we denote these runs Rogue, Tidal, and Rotating. Our
Rogue simulations are carried out in a (50Rp)3 box cen-
tered on the planet. The Tidal simulations use a similar
domain, but extended an additional 25Rp along the neg-
ative x -- axis towards the star ((75 × 50 × 50)Rp). Lastly
the Rotating simulations take the Tidal domain and ex-
tend an additional 25Rp along both the negative and
positive y -- axes ((75× 100× 50)Rp). For the planet's in-
ner boundary conditions, we use Rmask = 44/64Rp and
Rib = 49/64Rp. The edge of the planetary atmosphere
√
is at Re,0 = 89/80Rp, and the first ambient medium's
edge is at Re,1 = 2
3Rp (defined in Appendix B).
Computational expense varies greatly across our sev-
eral different simulations. We used 115 Intel Xeon E5-
2650 CPUs in parallel for all simulations. Wall times
ranged from roughly 2 days (Rogue) to a month (Rotat-
ing with a strong stellar wind). However, most simula-
tions ran in under 10 days, with the strong stellar wind
being the outlier (for the Rotating domain, the interme-
diate stellar wind took 13.5 days and the weak stellar
wind took 10 days). The two main reasons for the range
of cost were the varying domain sizes, and the inclusion
of a stellar wind. The most prohibiting factor, a strong-
stellar-wind, remains present on the finest level of reso-
lution throughout the duration of the simulation. As the
stellar wind is both hot and fast, the Courant condition
leads to smaller time-steps for the stellar wind than the
planetary outflow at the same spatial resolution. Ideally
in the future, adaptive mesh refinement or better static
mesh refinement will be chosen to avoid excessively high
spatial resolution in the stellar wind.
Resolution for the simulation is set such that within
a cell there is at most a scale height. While we do not
know the structure of the wind in steady state a pri-
ori, the hydrostatic isentropic atmosphere contains some
of the smallest scale heights throughout the duration of
the simulation. The number of scale heights between
two points in our hydrostatic isentropic atmosphere is
given by Eq. (C2). Since the hydrostatic atmosphere
and base of the wind require the most resolution, we use
a statically-refined mesh. The highest resolution region,
which is a box of size (2.5Rp)3 centered on the planet
with δx = Rp/64, encapsulates both the wind base and
hydrostatic atmosphere. We then include four coarser
levels of refinement around that, leading to a minimum
resolution δx = Rp/4 in the region outside a box of size
(4Rp)3 centered on the planet.
The star is modeled after the Sun, with M(cid:63) =
1.989 × 1033 g = M(cid:12) and R(cid:63) = 6.957 × 1010 cm = R(cid:12).
The ionizing flux is F∞ = 2 × 1013 cm−2 s−1, which is
comparable to our sun's "moderate to low solar activ-
ity" extreme-ultraviolet flux15 scaled to 0.05 au (Woods
15 Photon energies of 13.6 eV to 40 eV (91 nm to 30 nm).
10
McCann et al.
et al. 1998). Note our ionizing flux is scaled to the cen-
ter of our box as we use plane-parallel radiation. To
ramp the flux to avoid transients we use f0 = 10−2, and
t1/2 = 5 × 104 s. We use a monochromatic ionizing flux
of hν = 16 eV, which is in line with previous studies and
reasonable for a hot Jupiter around a quiet solar analog
(Tripathi et al. 2015).
For simulations that include a stellar wind, we con-
sider a range of wind strengths, as parameterized by
n(cid:63),0, Eq. (25). Our choices for n(cid:63),0 are 1.5 × 103 cm−3,
1.5 × 104 cm−3 and 7.0 × 104 cm−3 at reference radius
r(cid:63),0 = 4 × 1011 cm ≈ 0.03 au. Note that the proton num-
ber density of the Sun's stellar wind at this distance is
roughly n(cid:63),0 = 8.4 × 103 cm−3, which falls directly be-
tween our low and intermediate value. Vidotto & Bour-
rier's (2017) model of GJ 436 estimate a ram pressure
of Pram = 1.4 × 106 dyn cm−2 at the location of GJ 436
b, which nearly corresponds to the pressure of our inter-
mediate stellar wind. Vidotto et al. (2018) modeled HD
219134, and found pressures near HD 219314 b similar
to our strong stellar wind. The other parameters of the
stellar wind are T(cid:63) = 1.35 × 106 K and an initial veloc-
ity of v(cid:63) = 200 km s−1, corresponding to a Mach number
of M = 1.04. The stellar mass-loss rates correspond-
ing to our three wind strengths are 1.3 × 10−16 M(cid:12) yr−1,
1.3 × 10−15 M(cid:12) yr−1 and 5.9 × 10−15 M(cid:12) yr−1.
4. RESULTS
We now explore the effects of the tidal gravity, the
Coriolis force, and the stellar wind on the planetary out-
flow piece-by-piece. Our base case is an atmosphere re-
ceiving ionizing flux in the planet's potential with no
external stellar wind (Rogue run, § 4.1). The impact of
non-inertial forces from the planet's orbit around a star
is first examined without the Coriolis force (Tidal run,
§ 4.2) then with (Rotating run, § 4.3). The morphology
of the outflows until this point is recapped with a more
quantitative analysis of their velocity structures in § 4.4.
Next the effects of a stellar wind are demonstrated by
contrasting the results of varying stellar wind strengths
in both the Rogue and Tidal runs (§ 4.5). Finally the
full suite of stellar environment physics is considered in
the Rotating run with a stellar wind (§ 4.6). A sum-
mary of quantitative variables, particularly the mass-loss
rate, across all simulations is given in § 4.7, and obser-
vational consequences are delayed until § 5. Summarized
in Table 1 are the important simulation parameters used
across all runs.
4.1. Rogue simulation: effects of ionizing radiation
We begin with a planet receiving plane-parallel ioniz-
ing radiation in the absence of tidal gravity or a stellar
wind, shown in Figures 2(a) and 2(d). The closest phys-
ical analog would be an ejected rogue planet heated by
high-energy photons from a nearby high-mass star. Such
a situation would be more likely to occur in extreme en-
vironments, for example the young star-forming regions
near the Galactic Center. However, the motivation for
this study is as a base case for comparison to more com-
plex planetary conditions, i.e., those with a stellar host.
Despite inherent asymmetric heating, the outflow at
large scales is strikingly spherically symmetric. The sym-
metry arises from azimuthal pressure gradients freely re-
McCann et al.
Figure 2. The fully-launched density structure of a hydrodynamic escaping planetary outflow without a stellar wind. Ionizing radiation
enters from the left side of each box. The first row shows slices in the orbital plane, z = 0, while the second row shows vertical slices
perpendicular to the planet-star axis, y = 0. Left: Our Rogue planet, which only feel the effects of an external ionizing source and the
planet's gravity. Middle: Our Tidal simulation, which includes tidal gravity but no Coriolis term. Right: A true rotating frame with the
Coriolis force, called our Rotating simulation. This figure is available online as an animation, showing the launching of the outflow from
t = 0 s until t = 2 × 106 s.
distributing material along equipotential surfaces, the ef-
ficiency of which can be seen in the steady-state solution
of the temperature distribution, Figure 3. At larger dis-
tances material streams radially outwards, having sub-
dued the azimuthal pressure gradients. It is only on the
nightside, originating from the planet's shadow, that a
planetary tail and the surrounding dearth of material
break the symmetry. Note that the "planetary tail" is
not the nightside or downstream arm of the outflow,
but the noticeably neutral-enriched density enhancement
that forms in the shadow of the planet.
advects heat to the nightside, it cannot remain cold as it
does in the fixed anisotropic Parker wind.
Setting boundary conditions interior to the wind base
enables the nightside to readjust, preventing material
from reentering the atmosphere. Material advected to
the nightside, unable to receive ionizing radiation in the
planet's shadow, begins to form the planetary tail. The
tail then serves to lessen azimuthal pressure gradients
and shuts down the compression shock, explaining why
we only see the shock as an initial transient in our sim-
ulations. Even with typical ISM cooling rates at solar
metallicity, we find that the timescale for the material
to flow from the day to nightside would be too fast for
the tail to sufficiently cool and collapse. We note that
both the previous models and ours lack consideration of
conductive heating and cooling. While conduction was
found not to be dynamically important on the substel-
lar ray due to outflow timescales (Garc`ıa Munoz 2007;
Murray-Clay et al. 2009), if hot stagnant gas persists
on the nightside, evident in Figure 3, heat conduction
may be dynamically important -- something future multi-
dimensional simulations need to take into consideration.
The formation of the tail is also responsible for the sur-
rounding dearth of material. The tail grows by accruing
material via the tip, thus new material must travel far-
ther, spending more time in ionizing radiation, before it
can enter the shadow. Eventually there comes a point at
which this material has received enough ionizing energy
The planetary tail
is a new prediction in multi-
dimensional simulations (also seen in Debrecht et al. in
press), as previous models had not resolved the stellar
heating. Instead these models used an adiabatic index
near one, and fixed temperature boundary conditions
that varied as the angular distance from the substellar
point, to emulate an anisotropic, isothermal Parker wind
(Stone and Proga 2009; Carroll-Nellenback et al. 2017).
In place of a planetary tail, these prior studies show a
sustained nightside compression shock, which arises as
hot dayside material flows around the planet and enters
the planet's cold nightside atmosphere (See Figures 1 and
4 in Stone and Proga 2009; Figure 3 "ANISO" panel in
Carroll-Nellenback et al. 2017). We should note that this
phenomenon is also seen in our simulations, but only ini-
tially as our boundary conditions are set deeper than the
base of the wind. Thus, when inflowing dayside material
11
20020y (Rp)RogueTidalRotating20020x (Rp)20020z (Rp)20020x (Rp)20020x (Rp)10201019101810171016Density (g/cm3)McCann et al.
Figure 4. Panel (a): Streamlines (chartreuse yellow) overlaid on
gas density from the steady-state Tidal run (Figure 2(b)). The
sonic surface is demarcated in red, with the outer sonic surface
coinciding with the interface between the planetary outflow and
the virtually static ambient medium. Note that while supersonic
material does shock within the flow, the shock is oblique and thus
there is no corresponding sonic surface. Panel (b): Ballistic tra-
jectories (white) of particles launched radially from the surface of
the planet with velocity vp (see text). Ballistic particles launched
near the planet's terminator (x = 0, y = ±Rp) achieve the furthest
y -- extent for gas infalling onto the star. These trajectories bound
the rest of the flow and are in good agreement with the outer sonic
surface that demarcates the interface with the ambient medium.
Notice no significant trajectory crossing seems to occur in the or-
bital plane. Shocking results instead from trajectories crossing the
orbital plane from above and below, having been accelerated to
do so by stellar gravity -- white dots demarcate where a few such
trajectories, i.e., those originating with z (cid:54)= 0, pass through the
orbital plane. Trajectories launched near the poles (z = ±Rp) are
excluded, as they shock gas far out of the orbital plane.
ample flux to achieve such temperatures. We therefore
suggest a cosine anisotropic heating function would only
be suitable for planets whose dayside temperature is set
by the heating rate and not the cooling rate, such as a
planet similar to Earth.
4.2. Tidal simulation: effects of tidal gravity and
ionizing radiation
Our Tidal run is shown in Figures 2(b) and 2(e). In
addition to ionizing radiation, it includes tidal forces,
with the star located at x = −66.¯6Rp. Note that this
is not a true rotating frame as we neglect the Coriolis
force, but this frame is akin to that used in Tripathi et
al. (2015) and numerous 1-D models that included tidal
gravity.
Figure 3. The temperature of the gas for our Rogue planet. The
green contour is the surface where τ = 1 for ionizing radiation
entering from the left side of the plot. The corresponding dashed
green circle is of constant radius equal to the substellar τ = 1
radius. While the nightside is cold just at and below the dashed
circle, the temperature just above is practically uniform around the
entire planet. The pencil shadings are the line integral convolution
to assist in visualizing the flow. We can see a stagnation point in
the shadow near 4Rp, which leads to the flaring around the neutral
shadow exhibited in Figures 2(a) and 2(d). The sonic surface is
marked in red and does not form a closed surface.
to become unbound, and it flows radially outwards as
this is now the path of least resistance. Thus, drawing a
radial line from the origin to this point at the edge of the
shadow demarcates the dearth boundary. With this de-
scription the length, (cid:96), from the planet where the flaring
outflow begins to diverge from the tail can be estimated.
We equate the potential energy to the ionizing heating
rate multiplied by the time spent getting there, i.e., the
distance divided by the flow velocity, φ = G (cid:96)/v. For a
point-mass potential this yields
(cid:115)
(cid:96) =
GMp µv
Epe σHI F∞ e−τ .
(32)
With a flow velocity of order v = 10 km s−1, and
marginally bound gas near the τ = 1 surface, we find
(cid:96) = 3.6Rp -- close to the simulation's 4Rp. Naively one
might expect the tail to truncate at this distance as it
can no longer receive material at the tip. However, the
tail is hot and adiabatically expands outwards, allowing
new material to enter at points prior to the truncation,
sustaining its growth.
Lastly, we comment that irrespective of the night-
side steady state, a cosine temperature function does
not match the dayside temperature profile of the self-
consistently simulated hot Jupiter. At the altitude of the
wind base, the temperature is uniformly ∼7000 K, shown
in Figure 3. This directly calls into question whether
simple cosine anisotropic heating functions are valid in
the regime of hot Jupiters. We believe not, as the co-
sine function inherently assumes the temperature is set
by the flux received. Instead, we find that for the day-
side of hot Jupiters the temperature is set by the cooling
equilibrium temperature, as the entire face is bathed in
12
42024x (Rp)42024y (Rp)103104Temperature (K)McCann et al.
gravity. Trajectories originating both above and below
the orbital plane oscillate through the orbital plane, in-
tersecting many trajectories closer to the orbital plane
which were not as strongly accelerated into the plane.
Notice that with tidal gravity the dearth of material
around the tail appears less pronounced in the orbital
plane and nonexistent in the x -- z plane. The nonexis-
tence in the x -- z plane is due to the star's strong grav-
itational pull of all the gas back into the orbital plane.
Eventually far away from the planet, the entire night-
side stream will be compressed into a vertically pressure-
supported plane due to the star's gravity. In the orbital
plane the reduced flaring can be understood from the
tidal forces stretching the outflow along the x -- axis, i.e.,
material unbound from the planet no longer moves purely
radially. Unlike in the vertical plane, the dearth is not
readily filled by the stellar gravitational pull towards the
x -- axis, as the centrifugal force opposes the pull with a
push.
4.3. Rotating simulation: effects of tidal gravity and
ionizing radiation in a rotating frame
The Rotating run is presented in Figures 2(c) and 2(f),
where we have now included the Coriolis force in addi-
tion to the tidal forces. A Rotating frame without a
stellar wind has also been explored with a separate code,
AstroBEAR, finding qualitatively the same geometry and
quantitatively the same mass-loss rates (Debrecht et al.
in press). Due to the Coriolis force, streamlines now wind
in a clockwise fashion as our rotation vector points out of
the page (Figure 5(a)). We can still understand the day-
side and nightside arm dichotomy as owing to the tidal
focusing previously described, but now deflected with the
inclusion of the Coriolis force.
Including the Coriolis force in our ballistic trajectory
analysis demonstrates that there now exist significant
trajectory intersections, even in the orbital plane (Figure
5(b)). In Figure 5(b) the white dots along the ballistic
trajectories denote where the velocity of each stream-
line has been deflected by one radian, which defines our
Coriolis length scale. The surface created by these dots
is similar to the teardrop-shaped shock, evident in the
density map of Figure 2(c), with an additional rotation
of roughly π/12. This additional rotation may arise be-
cause the shocked gas moves slower than the unshocked
gas and therefore turns on a smaller scale. Material the
outflow shocks on has rotated more than predicted from
the ballistic analysis, modestly rotating the shock inter-
face everywhere.
Note that the Coriolis length is not spherically sym-
metric for two reasons. First, the velocity profiles along
each streamline may not be equal, owing to the day-
side/nightside asymmetric heating. This effect seems
small as our ballistic particles are all launched at the
same velocity from Rp and seem to give good agreement
with the simulation. However, a detailed examination of
the ballistic trajectories shows that the distance of the
nightside shock from the planet is not as well matched as
that on the dayside suggesting that the nightside would
be better described by modestly lower ballistic velocities.
The second and more significant asymmetry is due to the
fact that particles launched from the planet in quadrant
I (upper right) and III (lower left) are pulled by tidal
forces in in the same direction that the Coriolis force
13
launched with initial velocity vp =(cid:112)v2
Figure 5. Gas streamlines and ballistic particles launched from
the planet surface overlain on the steady-state density map from
our Rotating simulation (Figure 2(c)). The ballistic trajectories are
esc + v2∞ at Rp, which gives
the particles a total energy equal to that of the gas at infinity for
the Rogue case. Given this energy, the ballistic trajectories show
the farthest possible extent of the gas. The white dots along the
ballistic trajectories occur where the velocities have been deflected
by one radian from their original direction. This is what we call
our Coriolis length, which roughly matches the shock we see in the
simulation.
A clear difference between the Tidal and Rogue runs
is the funneling of material towards the star. Escaping
planetary gas is energetically restricted to a cone origi-
nating at the barycenter by the star's gravitational pull
and the frame's centrifugal push (Figure 4). This can
be further illuminated by considering ballistic particles.
Figure 4(b) illustrates the ballistic trajectories of parti-
cles launched radially from the planet's surface at veloc-
ity vp. For supersonic flow (external to the inner red
contour in Figure 4(a)) the internal energy of the fluid is
much less than the kinetic energy, and rough agreement
between the streamlines and ballistic trajectories is ex-
pected. Given that the outflow quickly becomes super-
sonic, large-scale streamlines should match appropriately
chosen ballistic trajectories. These trajectories are cho-
sen such that the total energy at infinity, ignoring work
done by tidal forces, is the same found in the Rogue runs,
e∞ = v2∞/2, so that vp =(cid:112)v2
esc + v2∞.
Along with the funneling comes a shock in the outflow
with the same funnel geometry. Examination of ballis-
tic particles in the orbital plane gives little hint of in-
tersecting trajectories (Figure 4(b)). However, due to
the lack of the centrifugal force in the z -- axis, intersec-
tions still readily occur. This can be seen by considering
ballistic particles launched with the majority of their ve-
locity parallel to the z -- axis (not shown). Trajectories
able to escape the planet's gravitational pull neverthe-
less fall back into the orbital plane due to the stellar
McCann et al.
4.4. Comparison of velocity structures
Figure 6 provides the x -- component (black) and the
y -- component (blue) of the velocity profiles for the
Rogue (dotted), Tidal (dashed) and Rotating (solid) runs
along the x -- axis.17 The Rogue planet has a velocity
x -- component that asymptotes to a finite value on the
dayside as expected from Eq. (19). On the nightside,
the column is roughly hydrostatic.18 For the Tidal run,
the velocity starts to significantly deviate from the Rogue
case outside the Hill radius (blue vertical). There is also
a hydrostatic tail on the nightside out to about 2Rp, af-
ter which tidal forces start significantly accelerating the
gas. Recall that the effective Hill sphere for the gas is
less than the Hill sphere in Eq. (26) due to gas pressure
forces (§ 2.4), so that the effective Hill radius appears to
be near 2Rp. This is almost equivalent to reducing the
planet's mass in Eq. (26) to Mp/3.
In a true rotating frame with the Coriolis force, stream-
lines are no longer radial and velocities along the x -- axis
no longer probe a single streamline. Looking at the
dayside, we can see the x -- component of velocity is less
than that of the Tidal run, as some of the velocity has
been transferred into the y -- component. Note that even
though the Coriolis force does no work, the velocity com-
ponents will not add in quadrature to be equivalent to
the Tidal run, as the same x -- coordinate corresponds to
different distances from the planet along a streamline in
the two runs. This explains why the nightside velocity
is greater in the Rotating case than in the Tidal case,
as the streamlines sampled here are originating closer to
the dayside and have had more time to accelerate. Note
that eventually both on the dayside and nightside there
is a shock even in the absence of a stellar wind. This oc-
curs near the Coriolis length scale where streamlines have
been significantly bent, and trajectories start to cross.
On the nightside the shock is at smaller distances for
reasons discussed in § 4.3.
4.5. Stellar wind in Rogue and Tidal runs:
effects of a stellar wind and ionizing radiation
In this section we add a stellar wind to both the Rogue
and Tidal simulations. We study three different stellar
wind strengths tuned by the stellar wind proton number
density, n(cid:63),0: a weak (n(cid:63),0 = 1.5 × 103 cm−3), an in-
termediate (n(cid:63),0 = 1.5 × 104 cm−3), and a strong stellar
wind (n(cid:63),0 = 7.0 × 104 cm−3). Qualitatively, what sepa-
rates the strong from the weak stellar wind is whether the
planetary outflow solution is a wind or a breeze. The in-
termediate stellar wind is in between these two extremes,
but is chosen to still be a wind but with a smaller bow
shock radius. Shown in the top row of Figure 7 is the x -- y
orbital plane slice of density in simulations without tidal
forces, i.e., Rogue runs with stellar wind. The second
row has tidal forces, i.e., Tidal runs with a stellar wind.
The first column is our weak stellar wind, the second our
intermediate and last our strong stellar wind.
The weak stellar wind (Figures 7(a) and 7(d)), is able
17 Along this cut, the x -- axis, the Cartesian variables directly
correspond to the polar variables: vx = vr and vy = vφ.
18 There is actually some infall coming from circulation in the
shadow as material falls inwards along the x -- axis and moves out-
wards near the edge of the shadow. See the line integral convolution
shading in Figure 3.
Figure 6. Velocity along the x -- axis for three simulations: Rogue
(dotted), Tidal (dashed) and Rotating (solid). The black lines are
the x velocity and the blue lines are the y velocity. In both the
Tidal and Rotating runs there is significant acceleration outside
the Hill radius from tidal forces. On the nightside the gas remains
stagnant out to about 2Rp,where tidal forces begin to accelerate
it away from the planet. The decrease in velocity in the Rotating
profiles at large separation is from a shock, which occurs near the
Coriolis length on the dayside, and slightly closer to the planet on
the nightside due to slower outflow velocities. The velocity on the
nightside is faster in the Rotating case compared to the Tidal case
because the x -- axis samples multiple streamlines in the Rotating
case. The verticals lines are clearly labeled, and overlap (or do
not exist for some simulations, i.e., the Coriolis length), except for
the sonic point. That is because due to tidal forces, Tidal and
Rotating (overlapping sonic points) go transonic before the Rogue
simulation. The shaded regions denote regions that are fixed by
boundary conditions (light grey for Rib and dark grey for Rmask),
and should be ignored.
bends them, towards the x -- axis (y = 0) and away from
the y -- axis (x = 0). Conversely, quadrants II (upper left)
and IV (lower right) are pulled in the opposite direction
of their deflections. Thus, in a rotating frame tidal forces
can either assist or hinder the turning of streamlines, and
the shock occurs closer to the planet in quadrants I and
III than in II and IV.
In Figure 5 the ballistic trajectories completely bound
the extent of the atmospheric escape. This is a statement
about which regions of space are energetically forbidden
for the escaping gas to reach. While the trajectories do
a good job in all four quadrants of probing the maximal
extent of the outflow, they dramatically overestimate the
extent of the outflow in quadrant IV. This is due to the
fact that the farthest reach of the flow to positive x val-
ues in quadrant IV comes from gas that originated or
passed through the shadow of the planet -- readily seen
by following the nightside tail in Figure 2(c). Unable
to receive ionization heating in the shadow, the gas is
not fully accelerated, turns on smaller lengths, and thus
reaches a lesser extent in quadrant IV than predicted by
ballistics.
The planetary tail remains present, but is now bent due
to the Coriolis force. Its turning length appears longer
than the other streamlines due to the column being virtu-
ally stagnant until the gas' effective Hill radius.16 Thus
the tail follows ballistic particles launched from the ef-
fective Hill sphere near 2Rp on the nightside, giving the
appearance of a longer turning length.
16 The effective Hill radius being smaller than the standard bal-
listic Hill radius, see § 2.4
14
-7.0-5.0-3.0-1.01.03.05.07.0x (Rp)40302010010203040Velocityx(km/s)40302010010203040Velocityy(km/s)40302010010203040Velocityy(km/s)40302010010203040Velocityy(km/s)RogueTidalRotatingSonic PointUV Depth OneHill RadiusCoriolisCoriolisHill RadiusMcCann et al.
Figure 7. Snapshots of the density structure of a fully launched hydrodynamic escaping planetary wind outflow with varying stellar wind
strengths in the orbital plane, z = 0. The first row shows our Rogue simulations and the second row is our Tidal runs. The columns are
for the various wind strengths parameterized by: n(cid:63),0 = 1.5 × 103 cm−3 (left), n(cid:63),0 = 1.5 × 104 cm−3 (middle), and n(cid:63),0 = 7.0 × 104 cm−3
(right). Panel (a) shows planetary gas forming a bow shock with the stellar wind. Panel (d) is unconfined as the stellar wind is not strong
enough to overcome the planetary outflow, and panel (e) is not in a steady state. The rest show complete confinement and breezes.
to confine the planetary wind when tidal forces are ab-
sent. In contrast when tidal gravity is included, the plan-
etary wind is no longer confined and is able to exit our
domain. The intermediate stellar wind (Figures 7(b) and
7(e)) is capable of confining the planetary wind for the
most part. Without tidal forces it is actually confined so
strongly that the planetary outflow is no longer a wind,
but a breeze as seen by the absence of shocked planetary
gas. With tidal forces the outflow is still capable of be-
ing a wind. However, we have provided a snapshot that
hints that the stellar confinement is not complete. This
is not due to picking a frame prior to steady state, but
as we shall see is due to no steady state standoff shock
being possible. Lastly our strong stellar wind (Figures
7(c) and 7(f)) is capable of confining both the Rogue and
Tidal simulations to breezes. Note that the Tidal simu-
lation is more confined, as within roughly 10Rp on the
nightside, the tidal forces are restorative to the x -- axis
(y = 0).
To explore the absence of a steady state in the inter-
mediate Tidal run we present a time series of simulation
snapshots in Figure 8. In Figure 8(a) the planetary wind
is reasonably well confined by the stellar wind. However
the confined region grows as the planetary wind contin-
ues to liberate more mass towards the star. Eventually
in Figure 8(b) the dayside outflow grows too close to the
star, where the stellar wind grows stronger, and is pushed
back. Finding the path of least resistance the overex-
tended gas is blown out and around the core-confined
planetary wind (Figure 8(c)), reducing the planetary
wind to a state similar to earlier, as seen in Figure 8(d).
Note that the time between Figure 8(a) and Figure 8(d)
is 2.7 × 105 s. This process repeats itself in perpetuity at
roughly this timescale. A movie of several cycles is avail-
able online. Since this is not a truly physical frame, we
leave a detailed explanation until § 4.6 -- where a similar
behavior is seen.
To understand the differences between a breeze and a
wind, we present a density comparison of the substel-
lar ray for three stellar wind strengths from our Rogue
run in Figure 9. Relative to when there is no stellar
wind (dotted line), the weak stellar wind (dashed line)
agrees well within ∼6 Rp, i.e., prior to the planetary gas
being shocked. The thickness of the planetary shock is
∼3 Rp, extending to the contact discontinuity between
the stellar and planetary winds at ∼9 Rp. By compar-
ison the intermediate stellar wind case (solid line) has
no shocked planetary wind, and the entire outflow is
in causal contact with the stellar wind as the outflow
never goes transonic and is a breeze. The zoomed insert
demonstrates that interior to the sonic point, the weak-
stellar-wind case is in excellent agreement with the no-
stellar-wind case, whereas the intermediate stellar wind
causes a planetary breeze that significantly diverges from
the base case. Taken in conjunction with the fact that
no information propagates upstream in a sonic flow, the
15
20020y (Rp)RogueWeakIntermediateStrong20020x (Rp)20020y (Rp)Tidal20020x (Rp)20020x (Rp)10201019101810171016Density (g/cm3)McCann et al.
Finally, we add our three stellar winds into our Ro-
tating simulation, shown in Figure 10. The top row are
the x -- y orbital plane and the second rows the x -- z plane
density slices. The columns are from left to right: a
weak, intermediate and strong stellar wind. Similar to
the Tidal simulations with a stellar wind, the three wind
strengths exhibit three different regimes. The weak stel-
lar wind (Figures 10(a) and 10(d)), is incapable of con-
fining the planetary outflow within our domain. For the
intermediate wind, we see a snapshot of the partially-
confined planetary wind, which as in the Tidal simula-
tion (§4.5), undergoes periodic disruption events. Lastly,
the strong stellar wind is capable of completely confin-
ing the planetary outflow and reducing it to a breeze.
Zoomed-in plots of the velocity magnitude, temperature,
neutral number density and neutral fraction for each sim-
ulation are shown in Figure 11. The columns from left to
right again correspond to weak, intermediate and strong
stellar winds.
The intermediate wind cannot achieve a steady-state
standoff shock. The time series for the Rotating run with
this behavior is displayed in Figure 12. Arranged from
Figures 12(a) to 12(d) in chronological order, are snap-
shots of the x -- y orbital plane gas density. Initially the
planetary outflow is confined, but continues to grow as
more gas is liberated from the planet. Eventually the
arm grows too large and is blown out, resetting the con-
fined planetary wind to something similar to an earlier
state, after which the process repeats.
To understand the timescale of the disruption event
consider the following. The geometry of the outflow
forms a torus around the star.
such that the dayside
arm is collimated into a cylinder, length L and cross-
sectional radius s. The mass enclosed in the cylinder
grows proportional to M (t) ∝ M t/2, where the factor of
one-half accounts for the two arms of the outflow. Let
the length of the cylinder grow at some average velocity
(cid:104)u(cid:105), i.e., L(t) ∝ (cid:104)u(cid:105)t. Then a cylinder that has just been
disrupted starts to grow according to, L(t) = (cid:104)u(cid:105)t and
M (t) = M t/2.
We imagine that the inflow "nozzle" growing the cylin-
der from the planetary outflow is sufficiently pressurized
as to not be disrupted or displaced. Thus, we say that
the cylinder is disrupted when the stellar wind has dis-
connected the cylinder from this inflow. Equivalently
stated, the cylinder has been radially displaced a dis-
tance ∆dcol. = 2s away from the star. The stellar wind
imparts a force on the column equal to the pressure times
the cross sectional area, A(t) = 2sL(t) = 2s(cid:104)u(cid:105)t, so that
Fwind(t) = A(t)P(cid:63),total(a) = 2s(cid:104)u(cid:105)tP(cid:63),total(a). Thus,
the acceleration of the cylinder from the stellar wind is
acol.(t) = Fwind(t)/M (t) = 4s(cid:104)u(cid:105)P(cid:63),total(a)/ M . Solving
for the time of disruption from the equation of motion,
∆dcol. = acol.t2
disrupt/2 = 2s, yields
(cid:115)
tdisrupt =
2 M
(cid:104)u(cid:105)P(cid:63),total(a)
.
(33)
2,
the
intermediate
From Table
has
M = 3.8 × 1010 g s−1.
From Table 1 P(cid:63),total(a) =
1.3 × 10−6 dyn cm−2. Approximating (cid:104)u(cid:105) = 25 km s−1
from Figure 11, we get tdisrupt ≈ 1.5 × 105 s. The
regime
Figure 8. Quasi-steady state time series of the Tidal simulations
for the intermediate stellar wind case. The wind starts confined
in panel (a) but continues to grow and extend outwards as seen
in panel (b). Eventually the column grows out too far and can no
longer maintain itself in the face of growing stellar wind strength.
In panel (c) we see the disruption of the column, returning back
to a confined state as seen in panel (d). The time lapsed between
panel (a) and (d) is 2.7 × 105 s. This process repeats continuously.
This figure is available online as an animation, showing several
periods of the quasi-steady state.
Figure 9. Total densities along the substellar ray for the Rogue
simulations with no stellar wind (dotted), a weak stellar wind
(dashed) and an intermediate stellar wind (solid). When a stellar
wind is present, shocks and standoffs between the planetary and
stellar wind occur. A several order of magnitude decrease in den-
sity occurs at the standoff location. Note that weak stellar winds
do not affect the solution below the sonic point (red vertical), while
a stronger stellar wind can alter the solution (see zoomed insert).
In fact strong winds force a planetary breeze solution without a
sonic point (lack of solid red vertical). The dark green vertical line
is the UV depth one, and the orange vertical line shows the sonic
surface of the stellar wind, corresponding to where it is shocked on
the planetary wind (present only for the intermediate wind within
the plotted region).
stellar wind does not affect the planetary outflow interior
to the sonic point unless it is strong enough to confine
the planetary wind to a breeze.
4.6. Stellar wind in Rotating run: effects of a full
stellar environment
16
20020y (Rp)4020020x (Rp)20020y (Rp)4020020x (Rp)10201019101810171016Density (g/cm3)-10.0-7.5-5.0-2.5x (Rp)102110201019101810171016Density (g/cm3)No WindWeak WindInter. WindSonic PointUV Depth OneStellarShockMcCann et al.
Table 2
Planetary mass-loss rates
Figure 10. Snapshots of the density structure of a fully-launched hydrodynamic, escaping planetary outflow with varying stellar wind
strengths in a full rotating frame. The first row shows slices in the orbital plane (z = 0) and the second row shows vertical slices
perpendicular to the planet -- star axis, y = 0. The columns are for various wind strengths parameterized by: n(cid:63),0 = 1.5 × 103 cm−3
(left), n(cid:63),0 = 1.5 × 104 cm−3 (middle), and n(cid:63),0 = 7.0 × 104 cm−3 (right). For the weakest wind, the arms are able to extend to the
boundary of our domain, implying a significant torus-like structure may be possible. For the intermediate-stellar-wind case, the gas is
still able to significantly extend outwards but is eventually confined. Lastly, the strongest wind is able to completely confine the outflow
into a cometary-like structure. This figure is available online as an animation, showing the evolution of the outflow when a stellar wind is
introduced at t = 5 × 105 s until t = 2 × 106 s.
time lapsed between Figure 12(a) and Figure 12(d) is
1.8 × 105 s, close to our order of magnitude estimate. We
can also plug the disruption time into the length at the
time of disruption and get L(tdisrupt) ≈ 30Rp, again in
good agreement with what can be seen in Figure 12(b).
Velocity and density profiles along the x -- axis are pro-
vided in Figure 13. The zoomed inset again shows sig-
nificant differences near the wind's sonic point between
planetary outflows that are winds (no and intermediate
stellar winds) and those that are breezes (strong stellar
wind). In the velocity profiles, the locations where the
x -- and y -- velocity profiles intersect at zero are the stag-
nation or standoff points of the outflow. The shocks can
also be seen where the x -- velocity magnitudes suddenly
decrease. Again all velocity profiles for planetary winds
are identical until they are shocked on their stellar wind
outside their sonic point, while the confined breeze solu-
tion differs substantially.
n(cid:63),0 (cm−3) Rogue (g s−1) Tidal (g s−1) Rotating (g s−1)
0.0
1.5×103
1.5×104
7.0×104
generate a polygonal mesh of our data, the normal of
which is given by the gradient. Then by projecting the
interpolated velocity into this normal and multiplying by
the interpolated density and area of a mesh element, we
get the mass-loss rate through that surface. Summing
the mass-loss rate over the entire triangulated sphere we
then arrive at the total mass-loss rate of the planet. We
perform the calculation at 0.8 Rp, 2 Rp, and 15 Rp and
find excellent agreement. Presented here are the time-
averaged results averaged over those three radii.
Note. -- The time-averaged mass-loss rate of the sustained
planetary outflow (1 × 106 s ≤ t ≤ 2 × 106 s).
4.7. Planetary mass-loss rates
For the four wind strengths studied here in the Rotat-
ing case, Table 2 list the mass-loss rates averaged over
the last 106 s of our simulation. To determine the mass-
loss rates we choose a sphere of radius larger than our
reset radius and calculate the flux of material through
the sphere. We use the marching cubes algorithm to
4.0×1010
3.9×1010
3.9×1010
3.6×1010
3.9×1010
3.8×1010
3.8×1010
3.5×1010
3.8×1010
3.8×1010
3.8×1010
3.4×1010
Notice that all rates for a given simulation are within
5 % of each other across all wind strengths except for the
strongest wind (n(cid:63),0 = 7.0 × 104 cm−3). In this case, the
outflow is a breeze and has a noticeably lower mass-loss
17
402002040y (Rp)WeakIntermediateStrong4020020x (Rp)20020z (Rp)4020020x (Rp)4020020x (Rp)10201019101810171016Density (g/cm3)McCann et al.
Figure 11. The velocity magnitude (top), temperature (top middle), neutral density (bottom middle) and neutral fraction (bottom) for
the Rotating simulation for three stellar wind strengths: weak (n(cid:63),0 = 1.5 × 103 cm−3, left), intermediate (n(cid:63),0 = 1.5 × 104 cm−3, middle)
and strong (n(cid:63),0 = 7.0 × 104 cm−3, right). Quivers in top row indicate direction of velocity, and the black circles demarcate Rp.
18
201001020y (Rp)WeakIntermediateStrong201001020y (Rp)201001020y (Rp)201001020x (Rp)201001020y (Rp)201001020x (Rp)201001020x (Rp)0102030405060v (km/s)103104T (K)103104105nH (cm3)103102101100Neutral FractionMcCann et al.
Table 3
Length scales in outflows
Rogue (Rp)
Tidal (Rp)
Rotating (Rp)
rτ =1
rsonic
1.02
1.97
1.02
1.81
1.02
1.81
Note. -- The substellar radii of various length scales.
M ∝ M−1
p F0. Furthermore, our mass-loss rate for the
Rotating simulation without a stellar wind finds good
agreement with Debrecht et al.
(in press). Here they
explored a similar planet and ionizing flux,19 finding an
M = 3.35 × 1010 g s−1.
For reference, in Table 3 we provide the time-averaged
sonic radius and the radius where optical depth to ion-
izing photons is one for our simulations. These either
did not change over the wind strengths, or were non-
existent such as the sonic radius for the breezes. The
only discernible difference is in the sonic radius between
the Rogue and Tidal or Rotating simulations. This dif-
ference results from the tidal forces helping to accelerate
the gas to supersonic velocities earlier than occurs in the
Rogue simulation.
5. OBSERVATIONAL CONSEQUENCES:
LYMAN-α TRANSITS AND SPECTRUM
To explore the observational consequences of the fea-
tures found in § 4, we perform synthetic observations of
the Rotating runs across three stellar wind strengths.
Through these synthetic observations we are able to
probe the spatial extent of escaping hydrogen via obscu-
ration maps, transit light curves, and spectra. A number
of lines have been seen in escaping planetary winds. We
focus our synthetic observations around Lyman-α as it is
the feature with the greatest number of high-confidence
detections. By comparing features seen in synthetic ob-
servations, we can search for unique signatures to obser-
vationally disentangle our different physical scenarios.
The synthetic observation procedure is discussed in
§ 5.1, and the results for our three scenarios are pre-
sented in § 5.2. We comment on the conditions in which
we expect observational signals to persist in § 5.3.
In
§ 5.4 we review current observations and tentatively sug-
gest which regimes they may fall into. Note that our
simulation is for a generic hot Jupiter, and we only seek
to provide context for features that have already been
observed and predict some we may yet to discover.
5.1. Procedure
To perform synthetic observations we begin by con-
structing an image plane defined by x = 0, which shares
its origin with the planet and is perpendicular to the
substellar ray between the planet and the star. We then
solve the equation of radiative transfer along rays that
emanate from the observer through each pixel in the im-
age plane. The observer is arbitrarily chosen to be one-
hundred parsecs away from the system's barycenter on
the opposite side of the image plane from the star, such
that rays within the domain are virtually parallel. This
19 Same mass and flux; however, different initial conditions led
to a different τ = 1 surface.
Figure 12. Time series of burping in a full-rotating frame for the
intermediate wind for the Rotating simulation. In panel (a), the
outflow is confined within a dayside arm. As the arm continues to
grow it get pushed outwards by the stellar wind as seen in panel
(b). Eventually the arm grows out too far and is completely blown
out from the system and detaches from the rest of the outflow as
seen in panel (c). Afterwards in panel (d), we can see that the
outflow is in a similar state to our initial panel and the process
will repeat. The time between panels (a) and (d) is 1.8 × 105 s.
Available online as an animation, showing several periods of the
quasi-steady state.
rate, consistently about 10 % lower than the average of
the other three wind strengths. This is in agreement with
the fact that the stellar wind is now in direct contact with
the planetary wind base, which adds additional pressure
that the planetary breeze must overcome.
These mass-loss rates are about an order of magnitude
lower than those found in the fiducial case of Tripathi
et al. 2015, who modeled an order of magnitude higher
ionizing flux for a planet twice as massive in a simula-
tion analogous to our Tidal simulation with n(cid:63),0 = 0.0.
M = 1.9 × 1011 g s−1, is in excel-
Their mass-loss rate of
lent agreement with ours when adjusted using the scaling
from Murray-Clay et al. (2009) for energy limited escape,
19
402002040y (Rp)20020x (Rp)402002040y (Rp)20020x (Rp)10201019101810171016Density (g/cm3)McCann et al.
Figure 13. Left: Density profiles for the no-stellar-wind, intermediate-stellar-wind and strong-stellar-wind cases of the Rotating simulation
(Figure 2(c) and Figures 10(b) and 10(c) respectively). From the zoomed inset we can see significant deviation in the density interior to the
sonic point in the high wind case. Unlike for weaker winds, the high wind case does not have a sonic surface (notice the lack of a solid red
line over plotted on the dashed and dotted lines), and has been confined to a breeze. Right: Velocity profiles for the same three simulations.
Where both the x -- and y -- velocities are zero is a standoff point between the planetary outflow and the stellar wind. Shocks in the velocity
profile can be seen as sudden changes to the velocity profile, near the Coriolis length for the no-stellar-wind case, and purely by coincidence
at the Hill Radius for the intermediate-stellar-wind case. The velocity profiles for the no-stellar-wind and intermediate-stellar-wind cases
are in excellent agreement prior to shocking.
produces an obscuration map appropriate for the mid-
dle of the planet's transit across the star. We produce
observations at various orbital phases by rotating the im-
age plane around the barycenter, Figure 14 (see § 2.8 of
Murray and Dermott (1999) for an example of a rotation
matrix in terms of Keplerian elements). As our simula-
tion is local, the orbital phases which we can observe are
limited by the box size. Yet more limiting is the extent
of the outflow in our domain, which in practice conser-
vatively limits the accessible orbital phases to φ = ±25◦.
We do not model the stellar Lyman-α spectrum or
interstellar extinction in this work. We instead calcu-
late the optical depth through the planetary outflow as a
function of frequency by integrating the absorption coef-
ficient along each ray, τν =(cid:82) αLyα(ν, s)ds. The absorp-
tion coefficient for Lyman-α is given by
(cid:16)
1 − e−hν/kBT (s)(cid:17)
πe2
mec
αLyα(ν, s) = nHI(s)
f12
φVoigt(ν, s).
(34)
Here e is the elementary charge, me the mass of an elec-
tron, c the speed of light, f12 = 0.4164 the absorption
oscillator strength between n = 1 and n = 2 for hydrogen
and φVoigt(ν, s) is the Voigt profile. Note that the term
in the parentheses is a correction to absorption for stim-
ulated emission. Since our simulations are already in the
star's rest frame, when calculating the Voigt line profile
we only need to shift the Doppler broadening component
by the projected bulk velocity of the fluid. We take for
granted that observers account for systematic velocities,
e.g., the relative velocity between the system and the ob-
server and orbital motion of the star, and neglect these
in our presentation.
From the optical depth, we calculate the stellar disc
averaged obscuration fraction
Figure 14. Schematic of viewing angle through the simulation
(bounded by the rectangle) in the orbital plane. The dashed lines
show the extent of the outflow transiting the star at a given orbital
phase. The half-circle centered on the star represents the orbit of
the planet around the star, with major ticks every 10◦ and minor
ticks every 5◦. While the image plane is defined on this circle,
rays extend through the entire domain by entering and exiting at
the boundaries. We only take our observations out to φ = ±25◦,
as for some simulations, larger angles do not intersect any of the
outflow. All distances and sizes are to scale. Refer to Figure 11 of
Carroll-Nellenback et al. (2017) for a global schematic.
for comparisons to observational spectra. Here A is the
area of the stellar disc defined as the surface S, dΣ is
the infinitesimal area element, I is the intensity observed
and I0 the intensity of light emitted from the star. Note
we only consider absorption without emission, such that
τν = ln(I0/I). We do not model the stellar line, nor
do we consider spatial variation across the stellar sur-
(cid:90)(cid:90)
(cid:18) I0 − I
(cid:19) dΣ
S
I0
A
(cid:104)Oν(cid:105) =
(cid:90)(cid:90)
(cid:0)1 − e−τν(cid:1) dΣ
A
S
, (35)
20
=
-10.0-7.5-5.0-2.5x (Rp)102110201019101810171016Density (g/cm3)No WindInter. WindStrong WindSonic PointUV Depth OneCoriolisStellarShockStellarShock-7.0-5.0-3.0-1.01.03.05.07.0x (Rp)40302010010203040Velocityx(km/s)40302010010203040Velocityy(km/s)40302010010203040Velocityy(km/s)40302010010203040Velocityy(km/s)No WindInter. WindStrong WindSonic PointUV Depth OneHill RadiusCoriolisCoriolisHill Radius=+25(t=+10.5 h)=0(t=0 h)=25(t=10.5 h)face, including limb effects. Typically obscuration will
be strongest near line center, where observers cannot
make accurate measurements due to geocoronal confu-
sion and interstellar extinction, and weakest at large ef-
fective Doppler velocities. Thus, a quoted transit ob-
scuration will depend on the frequency domain of the
spectrograph and methodology for removing geocoronal
confusion.
McCann et al.
For our synthetic observations we chose to measure
the obscuration between 1215.26 A to 1216.08 A (equiva-
lent Doppler velocities of −100 km s−1 to 100 km s−1),
ignoring obscuration between 1215.53 A to 1215.81 A
(−35 km s−1 to 35 km s−1) to model geocoronal confu-
sion and ISM absorption. Note that negative velocities
probe gas moving away from the star with blueshifted ab-
sorption, and conversely positive velocities correspond to
gas infalling towards the star with redshifted absorption.
Within this spectral range we further subdivide our ob-
servations into four unique frequency domains to retrieve
dimensionless equivalent widths for transit observations.
The dimensionless equivalent widths are calculated as
(cid:90) νb
νa
W =
(cid:90) νb
νa
(cid:104)Oν(cid:105) dν
∆ν
=
(cid:104)1 − e−τν(cid:105) dν
∆ν
.
(36)
Here νa and νb define the spectral range with a width
of ∆ν = νb − νa. Observations integrated over the red
range, 1215.81 A to 1216.06 A (35 km s−1 to 100 km s−1),
are referred to as "Red" and those over the blue range,
1215.26 A to 1215.53 A (−100 km s−1 to −35 km s−1), as
"Blue." The measurement called "Both" refers to an
integration over the entire observational range (excluding
the geocornal confusion region), which covers both the
red and blue wings.
The last measurement called "Full" is integrated over
the full observed spectrum as if there existed no geo-
coronal confusion or interstellar extinction. Typically
obscuration is largest near line-center and the "Full" line
displays the strongest obscuration. However, when the
bulk velocity significantly shifts the Doppler broadening
core away from line center, the dimensionless equivalent
width will be larger in the respective wing as the wings
are normalized over a smaller ∆ν. As the "Full" line
probes the total column density, it is suggestive of other
hydrogen lines, or metals with abundance proportional
to hydrogen, that have no confusion at line center.
As light rays take approximately 37.5 s to traverse the
domain's width of 75Rp, during which the fluid structure
does not significantly evolve, we are well justified in us-
ing the fast-light approximation, and produce synthetic
observations in post-processing. Less obviously justified
is using a single static output for the entire orbit when
performing our synthetic transits. Yet, there is only one
regime of stellar winds, the intermediate regime, in which
the simulation does not reach a steady-state. Using ap-
propriately timed outputs from the hydrodynamic simu-
lation would have no effect for systems in steady-state,
so the only concern would be for the intermediate stel-
lar wind regime. Yet, since the transit duration of the
planetary disc takes ∼4 h and the non-steady-state inter-
mediate regime has a periodicity of ∼50 h, it is acceptable
to use a static frame for the duration of a single transit
observation. As it is arbitrary when a transit might oc-
Figure 15. Upper Left: The transit (duration shown in vertical
dashed lines) for the weak stellar (Figure 10(a)) wind shows about a
1 % enhancement at observable frequencies (Red, Blue, Both) over
the planetary disc (dashed horizontal line). The Full line is for ob-
servations without geocoronal confusion and interstellar extinction,
telling us that most of the obscuration occurs in this domain. Up-
per Right: The spectrum is symmetric about line-center (vertical
dashed line) with some natural broadening enhancement over the
planetary disc (horizontal dashed line) for the red and blue regions
during transit (0◦ and 3◦). Out of transit (5◦) the only appreciable
absorption occurs from Doppler broadening in the geocoronal con-
fusion domain, which is unobservable. Lower Panels: Snapshots
at direct transit of the (50 × 20)Rp spatial obscuration at various
spectral frequencies denoted by their equivalent Doppler velocities.
The stellar disc is denoted by the red circle and the white bar pro-
vides the length scale at the plane of the planet. All line of sight
projected velocities are below 50 km s−1, with the fastest perhaps
detectable feature being the dayside arm at positive velocities (red-
shifted absorption) occurring before direct transit. Available online
as an animation, showing transit in the obscuration maps for the
Red, Blue and Both frequency domains, as well as a sweep of the
spectrum at φ = 0◦.
cur during the periodic disruptions, since the disruption
is not related to the orbital period but rather the stellar
ionizing flux and wind strength, we perform synthetic ob-
servations at the three different phases shown in Figures
12(a), 12(b), and 12(c).
5.2. Synthetic observations of planetary escape in
stellar environments
5.2.1. Weak-stellar-wind regime
With a weak stellar wind, observations are symmetric
both in the spectra and transit light curves, as seen in
Figure 15. The observable obscuration comes primarily
from the disc of the planet which occults 4.6 % of the stel-
lar disc during its transit between φ = ±4.9◦, or between
t = ±2 hours relative to direct transit. Since velocities
of the outflowing planetary wind are well within the geo-
21
201001020True Anomaly (°)102101100Equivalent widthFullRedBlueBoth-9-6-30369Time (hours)2001000100200Velocity (km/s)102101100Average obscured fraction=0°=3°=5°1215.01215.51216.01216.5Wavelength (Å)v = -150 km/s10 Rpv = -100 km/s10 Rpv = -50 km/s10 Rpv = -30 km/s10 Rpv = 0 km/s10 Rpv = 30 km/s10 Rpv = 50 km/s10 Rpv = 100 km/s10 Rpv = 150 km/s10 Rp103102101100Obscuration fractionWeak-stellar-wind synthetic observationscoronal confusion and interstellar extinction limits, the
outflow contributes a small enhancement (∼1 %) in the
"Red" and "Blue" spectral regions due to natural broad-
ening near mid-transit, when the absorption column is
largest.
Given that significant detections have been made far
outside the geocoronal confusion limits, yet our parame-
ters show no such features, it is worth asking how much
the outflow velocity would need to increase for our re-
sults to produce significant absorption in the "Red" and
"Blue" wings. To explore this question we consider an al-
ternative frequency domain from 1215.33 A to 1216.02 A
(−85 km s−1 to 85 km s−1), which we call our "Low"
domain, and ignore obscuration between 1215.59 A to
1215.75 A (−20 km s−1 to 20 km s−1). We note that our
weak stellar wind is the only scenario which does not
produce significant obscuration in our fiducial spectral
ranges, so we only consider the "Low" spectral range
here. This adjustment mimics making observations of
outflows boosted by 15 km s−1 in our standard spectral
domains, which we justify as follows.
Since natural broadening wings are not substantial in
our weak wind spectra, the line profile for the obscuration
is primarily due to Doppler broadening. Hence, where
absorption is significant, φν ∝ exp(−u2/b2), where u is
the velocity and b =(cid:112)2kT /µ is the Doppler broadening
parameter. Then integrating over our "Low" domain is
equivalent to integrating over our standard domain in a
wind of faster outflow velocities
e−u2/b2
du =
e−w2/b2
dw.
(37)
(cid:90) ub±s
ua±s
(cid:90) ub
ua
Here [ua, ub] is our fiducial range and s is the shifted
boost which relates our "Low" domain to our fiducial
range. Note that, w = u ∓ s with a top sign for the
red range and bottom sign for blue. Thus from Eq. (37),
we find that comparisons between our shifted "Low" ob-
servations and our standard observations approximately
corresponds to comparing to an outflow boosted by
15 km s−1. This is a first-order approximation as out-
flows with larger velocities may have different neutral
fractions and overall structure.
The transit measurement for the "Low" domain is pre-
sented in Figure 16. Compared to the fiducial observa-
tional range, the absorption is still symmetric at φ = 0◦,
while prior to φ = 0◦ there is substantially more red ab-
sorption. Conversely, roughly three hours after transit
blue is more dominant, though at a much lower ampli-
tude than red is prior to transit. This is due to which
arm of the outflow is being probed. The dayside arm
(redshift) moves faster than the nightside arm moving
away from the star (blueshift). Thus, a more significant
absorption feature can be seen in the red wings of the
spectra leading the direct transit (see the v = 30 km s−1
snapshot of Figure 15). Note that in the dayside arm,
some gas is still moving away from the star as it has
been turned by the Coriolis force (see Figure 11(a)). This
becomes more pronounced at larger distances from the
planet, and the "Blue" measurement eventually domi-
nates over "Red" prior to φ = −20◦.
5.2.2. Intermediate-stellar-wind regime
McCann et al.
Figure 16. Transit light curve of an outflow interacting with
a weak stellar wind, observed in our "Low" wavelength window
of 1215.33 A to 1216.02 A (−85 km s−1 to 85 km s−1). These ob-
servations, when compared with our standard window 1215.26 A
to 1216.08 A (−100 km s−1 to 100 km s−1), approximate planetary
outflows boosted by 15 km s−1. With larger outflow velocities, sig-
nificantly more obscuration occurs. Before transit there is signifi-
cant absorption at redder wavelengths due to the leading dayside
arm outflow infalling towards the star. Long before transit, blue
absorption is seen due to the Coriolis force turning the dayside arm
at large distances. The nightside outflow, probed after transit, is
slower and produces a weaker signal. Absorption is dominated
by the naturally-broadened line wings and is roughly symmetrical
between the red and blue wing.
For the intermediate-stellar-wind regime, we generated
synthetic observations from the static frames of Figures
12(a), 12(b), and 12(c). We found that though the
transit signal quantitatively differs between snapshots, it
qualitatively remains the same. Therefore, in Figure 17
we present the results from the Figure 12(b) snapshot,
which showed the strongest signal, to probe the gen-
eral features of the intermediate regime. As discussed
in § 5.1, the disruption timescale is ∼50 h and a tran-
sit is only ∼4 h. While using a static output is accurate
enough for transit, if we want to create accurate transit
light curves spanning ∼22 h (φ = [−25◦, 25◦]), we need to
perform observations at times consistent with the orbital
phases. Therefore, we also produce three time-resolved
transit light curves displayed in Figure 18, beginning at
φ = −25◦ for Figures 12(a), 12(b), and 12(c).
As in the weak stellar wind case, during the planetary
transit there is not much enhancement over the planetary
disc. However, unlike for the weak stellar wind, we find
a significant feature that precedes the planetary transit
within our standard frequency domain. As seen in the
spectra of Figure 17, when probing the dayside arm at
φ = −10◦ and φ = −20◦ the features are not from the
naturally-broadened wings of a Voigt profile, but from
the Doppler-broadened core. Therefore, while the stel-
lar wind may enhance the column density, the primary
cause of the increased obscuration is due to negatively-
accelerated material. In other words, the stellar wind is
22
201001020True Anomaly (°)10-210-1100Equivalent widthLow FullLow RedLow BlueLow Both-9-6-30369Time (hours)McCann et al.
Figure 17. Panels have the same layout as Figure 15, but are
now for our intermediate stellar wind (specifically Figure 12(b)).
Upper Left: Only ∼1 % enhancement over the planetary disc oc-
cultation occurs during transit, but prior to the transit strong blue
features are visible.Upper Right: No significant spectral asymme-
try is apparent during direct transit. Prior to transit, blue features
dominate. Lower Panels: Between 0 km s−1 to 50 km s−1 one can
clearly see the truncated dayside arm, while at large negative ve-
locities we can see that further upstream the extended disrupted
outflow causing significant obscuration. Available online as an ani-
mation, showing transit in the obscuration maps for the Red, Blue
and Both frequency domains, as well as a sweep of the spectrum
at φ = 0◦.
pushing the outflowing gas towards the observer.
The strong blue-shifted absorption signal is strongest,
∼20 % for Figure 12(b), but is present in all three snap-
shots in time. The major difference is the amplitude of
the signal, which is weakest for Figure 12(c) at ∼5 %.
Comparisons between Figures 12(b) and 12(c) show that
these two phases have the most and least amount of
dense gas being blown out by the stellar wind. Whether
the pre-transit signal ever disappears depends on the
timescale for the stellar wind to disperse the dayside ma-
terial, and the timescale for another disruption to occur.
For our parameters, neutral material outwardly acceler-
ated by the stellar wind is always around, and the signal
is continuously present but modulated in amplitude.
In Figure 18 we time-resolve the transit (in ∆φ = 2◦ in-
crements) by ray-tracing through a series of appropriate
simulation snapshots. We chose three times that corre-
spond to φ = −25◦, thereby measuring the transit at var-
ious disruption stages (those being Figure 12(a), 12(b),
and 12(c)). Generally, the signal seen mimics two bodies
transiting the star in quick succession -- a double transit.
The first "object" (which is just the outwardly acceler-
ated dayside outflow) enters transit tens of hours prior
to the true transit of the planetary body. This would be
Figure 18. Three time-resolved transit observations of the unre-
solved Lyman-α line (our Both domain) for the intermediate stel-
lar wind. Each transit is labeled by the time corresponding to
φ = −25◦, which in time order correspond to Figure 12(a), 12(b),
and 12(c). Given the non-static disrupting dayside outflow, the
pre-transit signal varies throughout the various stages of the dis-
ruption. The signal typically appears to resemble a "double tran-
sit," in which a comparable signal to the transiting body appears
and disappears prior to the transit of the planet. We note that this
feature is strong when only the blue wing of the line is consider (c.f.
Figure 17).
easy to disentangle with a spectrally resolved observation
as it is entirely biased bluewards of line center; however,
as seen in Figure 18, an unresolved observation could
produce such "double transits." Additionally, an opti-
cal transit observation would not find any such "Trojan
body," indicating it was instead neutral hydrogen from
the dayside outflow.
5.2.3. Strong-stellar-wind regime
For our strongest stellar wind we observe a highly
confined cometary tail-like outflow (Figures 10(c) and
10(f)). This structure leads to a transit light curve that
is strongly asymmetric both temporally and in its spectra
(Figure 19).
The blue features are due to the stellar wind acceler-
ating the planetary outflow away from the star, so that
neutral gas asymmetrically absorbs more efficiently at
blueshifted velocities. Moreover, during transit substan-
tial obscuration occurs at large velocities due to Doppler
broadening skewed towards negative velocities. Shortly
following the transit, blue obscuration starts increasing.
This is the result of a continuously accelerated nightside
arm being blown out at faster and faster velocities by the
stellar wind, shifting the Doppler core to bluer frequen-
cies. Notice that in the spectra at φ = 0◦, the star is not
entirely obscured due to the confinement of the dayside
arm, visible in the obscuration maps of Figure 19, and
maximal obscuration is not obtained until later.
We now ask whether this signal
is due to a den-
sity enhancement by stellar wind confinement. An en-
hanced column density might increase absorption in the
23
201001020True Anomaly (°)102101100Equivalent widthFullRedBlueBoth-9-6-30369Time (hours)2001000100200Velocity (km/s)102101100Average obscured fraction=0°=10°=20°1215.01215.51216.01216.5Wavelength (Å)v = -150 km/s10 Rpv = -100 km/s10 Rpv = -50 km/s10 Rpv = -30 km/s10 Rpv = 0 km/s10 Rpv = 30 km/s10 Rpv = 50 km/s10 Rpv = 100 km/s10 Rpv = 150 km/s10 Rp103102101100Obscuration fractionIntermediate-stellar-wind synthetic observations201001020True Anomaly (°)102101100Equivalent widtht=1.64×106 st=1.70×106 st=1.76×106 s-9-6-30369Time (hours)McCann et al.
Figure 19. Panels have the same layout as Figure 15, but are now
for our strong stellar wind (Figure 10(c)). Upper Left: The tran-
sit in the strong regime has significant blue enhancement over the
planetary disc occultation during transit. Additionally, after the
transit strong blue features are visible. Upper Right: The asym-
metry during direct transit favoring blue absorption comes from
gas moving away from the star at the outer edge of the planetary
outflow. Conversely, after transit there are significant blue fea-
tures due to the entire outflow being accelerated outwards. Lower
Panels: The sharp confinement due to the stellar wind is easily
visible in most panels. At large negative velocities the effects of
the outer edge of the outflow being more strongly accelerated are
visible. Available online as an animation, showing transit in the
obscuration maps for the Red, Blue and Both frequency domains,
as well as a sweep of the spectrum at φ = 0◦.
naturally-broadened line wings. To understand the col-
umn density enhancement one obtains from stellar wind
confinement we make comparisons to the Rotating sim-
ulation with no stellar wind. This has been done previ-
ously in 2-D (c.f. Figure 8 of Stone and Proga 2009) and
is reproduced for 3-D simulations in a rotating frame in
Figure 20. We plot the column density along rays par-
allel to the x -- axis in the orbital plane as a function of
the impact parameter measured from the center of the
planet.20
To make comparisons to models not done in a proper
rotating frame, we have overlain the column density for
our Rogue simulation, which has a roughly spherical out-
flow. In the absence of a stellar wind (Rotating) the out-
flow is confined to a torus around the star due to tidal
forces. When a strong stellar wind is present (Strong
Rotating) the outflow's shape is still toroidal due to the
tidal forces, but is further constricted and thus enhanced
by the stellar wind.
20 Here we chose ray trace as a function of impact parameters,
rather than as a function of orbital phase, for direct comparison to
Stone and Proga (2009).
Figure 20. Column densities integrated through outflows with
spherical (Rogue), toroidal (Rotating), and constricted toroidal
(Rotating with a strong stellar wind -- Strong Rotating) geometries.
For small impact parameters the column densities in all geometries
match, as the outflow still behaves as if it is spherically symmet-
ric until tidal forces or the stellar wind can significantly alter the
flow into a torus. While the column density in the torus geometry
should be roughly constant, notice that for the leading arm of the
torus the density actually increases around 10Rp. By referring to
Figures 10(a), we can see this density enhancement is due to the
shock at the Coriolis length.
Notice that outflow confined to a torus significantly
enhances the column density over spherical outflows at
large impact parameters. This can be understood by
considering the area through which the planetary mass
loss flux passes in each case. For a spherical outflow, this
fluxing area is a sphere, which scales as r2, while for the
torus the fluxing area is the constant-area end caps of a
cylinder. Thus, as long as the velocity does not change
and the outflow has reached ionization equilibrium, the
neutral column density for a torus will remain constant.
Given a fixed mass-loss rate, stellar wind constriction
of the torus leads to a column density enhancement.
This can be modeled by considering a cylinder of mass
M , length L, and cross sectional radius r, radially con-
stricted to a cylinder of length L and radius s. While the
column through the cylinder, 2r to 2s, decreases by a fac-
tor of s/r, the density increases by a factor of (r/s)2, so
that column density is enhanced by r/s. Our strong stel-
lar wind constricts the cross sectional radius roughly by
half, seen by comparing Figures 10(d) and 10(f). While
the mass-loss rate of a breeze is lower than the corre-
sponding wind (Table 2), the difference is within ten per-
cent for our parameters. Hence, relative to the weaker
stellar wind cases, for our strong-stellar-wind planetary
breeze, constriction increases the neutral density in the
nightside arm by almost a factor of two. Another fac-
tor of two comes from redirecting the escaping gas from
the dayside arm into the nightside. Thus, we estimate
a grand total enhancement of a factor of four. This es-
timate matches what we see in the nightside arm of the
constricted torus, between −30 Rp ≤ b ≤ −10 Rp, com-
pared to the unconstricted torus (Figure 20).
We conclude from Figure 20 that the stellar wind does
24
201001020True Anomaly (°)102101100Equivalent widthFullRedBlueBoth-9-6-30369Time (hours)2001000100200Velocity (km/s)102101100Average obscured fraction=0°=10°=20°1215.01215.51216.01216.5Wavelength (Å)v = -150 km/s10 Rpv = -100 km/s10 Rpv = -50 km/s10 Rpv = -30 km/s10 Rpv = 0 km/s10 Rpv = 30 km/s10 Rpv = 50 km/s10 Rpv = 100 km/s10 Rpv = 150 km/s10 Rp103102101100Obscuration fractionStrong-stellar-wind synthetic observations402002040Impact Parameter (Rp)101310141015101610171018Neutral Column Density (cm2)RogueRotatingStrongRotatingsubstantially enhance the neutral column density over
that from a spherical outflow. However, tidal gravity and
the Coriolis force -- by themselves confining the outflow
to a torus -- can also produce substantial enhancement in
the column density at large distances from the planet.
In fact, for our parameters the strong stellar wind only
increases the optical depth in the naturally-broadened
wings by a factor of a few over the weak-stellar-wind case.
This may seem contrary to previous work that suggested
stellar wind confinement could greatly enhance the tran-
sit signal, but the enhancement found in 2-D models
with a stellar wind and no Coriolis force is similarly due
to the geometry of constricting the flow to a torus (see
Stone and Proga 2009). Therefore, while models that ne-
glect the Coriolis force will see significant enhancement
at large impact parameters due to a stellar wind, those
that are done in a co-rotating frame shall not (unless
the stellar wind strength is much larger compared to the
tidal gravity than simulated here). We further note that
our measured enhancement is modest at small impact
parameters (b (cid:46) 10Rp), where most observations have
so far probed. The lack of enhancement is due to the
fact that the outflow will still behave roughly spherically
until the outflow has been significantly altered by the
tidal and Coriolis forces -- becoming toroidal.
Lastly, and most importantly, an increased obscura-
tion coming solely from naturally-broadened line wings
from a low-velocity column density should be symmet-
ric. In our simulations the predominant effect of a stel-
lar wind is an asymmetric signal, as gas is exclusively
accelerated away from the star. Therefore, symmetric
observational features are likely not due to stellar wind
confinement, which can be confirmed from the lack of an
enhancement in the "Red" measurement of our strong-
stellar-wind regime (Figure 19) relative to other stellar
wind regimes.
5.3. Absorption outside direct transit
A prediction of this work is that planetary winds
are capable of significant obscuration at large distances
from the planet. For our parameters, the ionization
timescale is τion = (σHIF )−1 = 3.6 h. Generously av-
eraging the outflow velocity to 40 km s−1, we calculate
the gas has undergone nearly seven ionization timescales
by φ = ±20◦.21 Thus, the existence of large scale signals
from ionized gas may be surprising. However, note that
after many ionizing timescales a gas does not become
completely ionized; rather, it reaches ionization equilib-
rium set by the detailed balancing of ionization and re-
combination. Indeed for our simulations we have verified
that the ionization fraction is within a few percent of
equilibrium past 10 Rp.
Apparent then by our synthetic observations in § 5.2,
gas in ionization equilibrium is sufficient to produce
large-scale observable signals. However, these simula-
tions were performed for a low ionizing flux analogous to
the quiet Sun. We now ask if signals should persist at
higher ionizing fluxes by considering each contribution
to the optical depth. To order of magnitude the optical
depth is τν = (1− X)NH σLyα(ν), where X is the ioniza-
21 Note the time relative to direct transit is not equivalent to
the time it takes the gas to reach the location probed at the time
relative to direct transit.
McCann et al.
tion fraction, NH is the total column density of hydrogen,
and σLyα(ν) is the cross section of Lyman-α absorption.
The impact of the outflow structure on σLyα(ν) is already
discussed in § 5.2.
The total column density depends both on the to-
tal number density of hydrogen, nH,22 and the column
length, L. The column length scales with the Coriolis
length, which depends on F only through the velocity
of the planetary outflow. We make the assumption that
this dependence is weak and let L be independent of F .
From Murray-Clay et al.
(2009) we consider the two
regimes of escape: energy-limited and radiation-limited.
As more energy deposited into the atmosphere liberates
more mass, in the energy-limited regime M ∝ F and for
the radiation-limited regime M ∝ F 1/2.23 Since the cross
section of the tidal torus only depends on L, which we
keep constant, then M ∝ nH. Therefore, in the energy-
limited regime NH = nHL ∝ F and in the radiation-
limited regime NH = nHL ∝ F 1/2
With larger F one might suspect lower neutral frac-
tions (1 − X), as the ionization rate increases. As the
outflow at large distance is in ionization equilibrium
((1 − X)σHIF = αBnHX 2), the ionization fraction as
a function of flux
(cid:18)(cid:114)
1 +
− 1
(cid:19)
(cid:19)2(cid:33)
4αBnH
σHIF
(cid:32)(cid:18) αBnH
σHIF
X =
1
2
σHIF
αBnH
and
1 − X =
αBnH
σHIF
+ O
,
.
(38)
(39)
Here we have Taylor expanded in αBnH/(σHIF ) as we are
already in the ionized regime far in the outflow such that
(1−X) (cid:28) 1. Then using the scalings for n, in the energy-
limited regime, the neutral fraction does not change,
(1−X) is independent of F , while in the radiation-limited
regime the neutral fraction scales as (1−X) ∝ F −1/2. So
while there are more ionizing photons at higher fluxes,
there is also a higher number density to facilitate recom-
bination, preventing the neutral fraction from decreasing
rapidly with increasing flux.
Taken all together τν ∝ F in the energy-limited regime
and τν is independent of F in the radiation-limited
regime. This suggests that as fluxes increase, the sig-
nal increases until the radiation-limited regime, at which
point the signal is saturated. Thus in systems similar to
the ones simulated, these signals may indeed be robust.
5.4. Atmospheric escape observations
As seen in § 5.2 it is possible to differentiate between
our modeled stellar environments. Weak stellar winds
produce the most symmetric transits and spectra. Par-
ticularly fast planetary winds can produce pre-transit
absorption that is skewed red due to the dayside arm
infalling towards the star. Intermediate stellar winds are
22 Note that the total number density n = nH + ne (cid:54)= nH, where
nH = nHI + nHII is the total number density of the hydrogen
species.
23 In the radiation-limited regime more energy is lost through
radiative processes and escape becomes less efficient.
25
marked by a quasi-static disrupted outflow, which man-
ifests as time-varying blueshifted absorbing gas present
significantly prior to transit. Lastly, strong stellar winds
have strong spectral asymmetry during transit, little pre-
transit obscuration, and increasing blue absorption post-
transit from the strongly collimated tail being blown out
by the stellar wind. Thus, to accurately distinguish be-
tween stellar regimes, measurements both in and far out
of transit are required. Note that even a few hours out
of transit may be too close to transit, as it takes time for
the stellar environment to accelerate the outflow. Unfor-
tunately, since Lyman-α observations must be performed
by expensive space-based telescopes, observations far out
of transit are lacking.
A general theme amongst comparisons to observations
is that our predicted Lyman-α obscuration fractions are
smaller than what is observed. While our planetary pa-
rameters are marginally a hot Jupiter analog, there is
no one-to-one correspondence with any observed system.
Yet, this is unlikely to account for the entire difference
between the observations and our simulations, as previ-
ous modeling of specific systems also struggle to match
observations, e.g., HD 209458 b in Murray-Clay et al.
(2009). Since outflow velocities retrieved in similar mod-
els are consistently too slow to explain the observations,
it is widely thought that more exotic physics is needed
(charge exchange, magnetic fields, etc.). However, we
note that Eq. (19) implies that the asymptotic velocity
u∞ depends on the energy spectrum of the ionizing pho-
tons. A harder spectrum produces more energy per ion-
ization generating larger ∆q∞ and thus larger u∞ (most
easily conceptualized in the energy-limited regime, where
radiative losses are minimal). Whether a more complete
investigation of ionizing spectra can produce larger wind
velocities merits investigation. To explore the possible
consequences, we can lower the geocoronal confusion lim-
its in post-processing to mimic faster outflows to zeroth
order (see "Low" spectral frequency in § 5.2.1).
To compare our models to observations we consider
four different planetary systems, each of which has a
distinctive signature. The first observed exoplanet un-
dergoing escape, HD 209458 b, has a transiting decre-
ment of (15 ± 4) % in Lyman-α flux from 1215.15 A to
1216.1 A (Vidal-Madjar et al. 2004).24 The spectrum
was originally reported to be marginally asymmetric
(Vidal-Madjar et al. 2003), but further measurements
have not detected any significant asymmetry (Ben-Jaffel
2007). Since strong stellar winds significantly skew the
signal bluewards, even during direct transit, we rule out
the strong stellar wind regime. Without more measure-
ments out of transit it is hard to distinguish between a
weak and intermediate stellar wind. Our "Low" mea-
surements, which decreased the geocoronal confusion by
15 km s−1, produced a roughly 15 % symmetric signal
during transit. This could suggest that HD 209458 b
might orbit a star with a relatively weak wind, and a hard
ionizing spectrum. Alternatively, all absorption from this
system may result from physics, such as charge exchange,
not modeled here. In particular, symmetric enhancement
of the naturally-broadened line wings due to stellar wind
24 Two other frequency bin decrements often cited: (8.9 ± 2.1) %
from 1214.83 A to 1216.43 A, or (5 ± 2) % over the entire line
1210 A to 1220 A (Vidal-Madjar et al. 2008).
McCann et al.
confinement is not supported by our results.
Next, transits from HD 189733 b, with optical occulta-
tion of 2.4 %, found Lyman-α decrements of (5.0 ± 1.3) %
for the entire line, (14.4 ± 6.6) % for the blue wing be-
tween −230 km s−1 to −140 km s−1, and (7.7 ± 2.7) % for
the red wing between 60 km s−1 to 110 km s−1 (Bourrier
et al. 2013). We note that the observations have also
seen significant variability in Lyman-α absorption over
various epochs. However, this time variability is not in-
dicative of an intermediate stellar wind, as the variability
is during direct transit and not further out in the out-
flow. Rather, this implies that the mass-loss rate may be
periodically modulated, possibly by a spatially variable
stellar wind (Vidotto et al. 2018) or a temporally flar-
ing star -- consequences not probed by our steady stel-
lar winds and ionizing flux. We note that the numeri-
cal agreement with our strong-stellar-wind simulation for
blue absorption during transit is coincidental because our
modeled parameters match neither HD 189733 b nor the
wavelength window of the observations. Furthermore,
the asymmetric wavelength windows for these observa-
tions make evaluating the red/blue asymmetry during
transit challenging. Nevertheless we note that the red-
wing absorption is enhanced over the optical, a feature
not seen in our models except in our "Low" measure-
ments. Moreover, our blue frequency range is nearly
100 km s−1 slower than the blue range probed here, sug-
gesting that the blue-shifted material may have been ac-
celerated by a much stronger, and possibly faster, stellar
wind than what we have simulated.
(2015)).
The hot Neptune GJ 436 b has the strongest and most
asymmetric observations of any observed transit. The
transit is within the measurement error of being unde-
tectable in the red wings except in post-transit, with an
obscuration of (8.0 ± 3.1) %. For the blue wings the ob-
scuration is significant ∼2 h before, (17.6 ± 5.2) %, dur-
ing transit, (56.2 ± 3.6) %, and ∼2 h after the transit,
(47.2 ± 4.1) % (Ehrenreich et al.
It should
be noted that due to lower surface gravity making
the liberation of mass easier, atmospheric escape from
hot Neptunes will likely be more dramatic than from
hot Jupiters.25 Thus, while the observations roughly
correspond to velocities seen in our strong-stellar-wind
regime, which shows significant blue obscuration from
the cometary-like tail being blown outwards by the stellar
wind, GJ 436 b's wind is able to persist at larger orbital
phases on the dayside. As observations see blue absorp-
tion before transit the bow shock is not as close to the
planet as in our simulation, which could be accomplished
with a slightly weaker stellar wind. Conversely, consider
if the outflow disruption seen in the intermediate regime
occurs at the Coriolis length. This is plausible since at
the Coriolis length the velocity has deflected away from
the star, leading the planetary pressure support at the
stellar wind interface to be completely thermal -- even
if the outflow is supersonic (ram pressure dominated).
Since GJ 436 b is a Neptune, its outflow velocities may
be larger leading to a larger Coriolis length. Therefore, a
persistent leading large-scale observational feature could
be consistent with the intermediate-stellar-wind regime.
25 For Neptune g(cid:91) = ∼1100 cm s−2 v.s. g(cid:88) = ∼2730 cm s−2 for
Jupiter, where (cid:91) represents Neptune.
26
Bourrier et al. (2016) suggest that a key component
of GJ 436 b's observations can be explained by the ac-
celeration of planetary gas by radiation pressure. Their
model and the model reported here differ in several key
respects. First, because the model in Bourrier et al.
(2016) is a particle simulation, it does not include forces
from the pressure gradient of the gas. As exemplified in
the work reported here, pressure gradients cause gas to
be launched from the planet's surface and accelerated to
a non-zero bulk velocity away from the planet. Motivated
by the fact that this launch velocity is much smaller than
the large velocities probed in the line wings of Lyman-α
observations, Bourrier et al. (2016) use as their starting
condition a population of particles at ∼3 planetary radii
with thermal velocities but no bulk velocity. Because
this population lies within the planet's Hill radius, these
particles cannot escape until they are substantially accel-
erated either by radiation pressure or by the stellar wind,
both of which are modeled through probabilistic interac-
tions with the gas particles, including self-shielding.
These choices have subtle but important consequences.
First, this leads to a highly ionized planetary outflow
(relative to our simulations), as the "static" particles are
likely to be ionized outside τ = 1 while they wait to
be accelerated (something they note in §3.3.3). In prac-
tice, what this means is that the outflow becomes ionized
enough to be optically thin to radiation pressure before
it is accelerated out of the planet's potential well.
In
contrast, we find in our simulations that the planetary
outflow is extremely optically thick out to more than
10 planetary radii, well beyond the planet's Hill sphere.
The difference comes from the outflow velocity driven by
gas pressure forces -- though small, this velocity is large
enough to allow the planetary wind to leave the planet's
Hill radius before reaching ionization equilibrium.
Recall that optically thin radiation (τν (cid:28) 1) will only
absorb a tiny fraction of the incoming flux by defini-
tion, yet will do so throughout the entire outflow.
In
contrast, except at very large separations, we find that
the planetary wind is predominately optically thick to
the entire Lyman-α photon band. Thus, significant op-
tically thin radiation pressure (not modeled) is not an
important accelerant in the majority of our simulation
box, and certainly in the regions dominating our syn-
thetic observations. Optically thick radiation pressure
(also not modeled), in contrast, may impact the velocity
structure at the boundaries of our simulated outflow, ex-
erting a force at the wind interface rather than uniformly
throughout the outflow. As mentioned in § 6, optically
thick radiation pressure may work in tandem with stellar
wind pressure to shape the large scale structure and ve-
locities of the planetary outflow, but it does not provide
significant acceleration of the gas in the vicinity of the
planet. This contrasts with Bourrier et al. (2016) who
find that radiation pressure was found to be significant
part of the acceleration of the outflow, a result that may
not be surprising given that the model is accurate only
after the particles have been significantly accelerated by
the stellar wind or radiation pressure.
Additionally, particle models that do not model inter-
particle interactions cannot generate pressure forces,
which is an important part of the impact of the stellar
wind and optically thick radiation pressure as the force
McCann et al.
is distributed throughout the outflow.26 The impact of
both optically thin and optically thick radiation pressure
far from the planet merits further investigation in hydro-
dynamic simulations.
Lastly, 55 Cnc b does not have an optical transit de-
tection, yet intriguingly a (7.5 ± 1.8) % decrement be-
tween −76.5 km s−1 to 0 km s−1 was measured in the blue
wings of Lyman-α at its inferior conjunction -- the loca-
tion where a transit would occur if the planet was copla-
nar (Ehrenreich et al. 2012). No detectable obscuration
in the red wings was reported. This obscuration is in-
terpreted as coming from the transit of a fraction of an
extended hydrogen atmosphere. The spectral asymmetry
in the observation suggestively indicates a strong stellar
wind.
6. CONCLUSION AND FUTURE WORK
In summary, the stellar environment can play a signif-
icant role in shaping planetary outflows. Through our
bottom-up approach we have examined several compo-
nents of the stellar environment in 3-D: ionizing radia-
tion, tidal gravity, the Coriolis force and a stellar wind.
While spatially resolved observations of these outflows
are not feasible, we demonstrated that spectrally and
temporally resolved transit observations may still illumi-
nate the overall structure of the outflow.
Alterations to the planetary outflow considered here
come from orbital effects and the interaction with a stel-
lar wind. Tidal gravity and the Coriolis force funnel the
outflow into a torus with a dayside and nightside arm.
The toroidal geometry enhances the column density at
large distances over spherical outflows. Additionally the
inclusion of non-inertial forces lead the outflow to shock
on itself, justified by analysis of ballistic particle trajec-
tory crossings. For stellar winds we find three unique
regimes with respect to their effects on the planetary
outflow. From our study of stellar environments and the
morphology of planetary outflows we summarize the fol-
lowing key points
1. Whether the planetary outflow is a wind or a
breeze, depends in part on the stellar environment.
However, if the the planetary outflow is a wind,
then the mass-loss rate is insensitive to the stellar
environment.
2. Large detectable signals may be present far outside
of transit.
3. These large scale signals help probe the stellar en-
vironment
i. Weak stellar winds cannot restrain the plane-
tary outflow, resulting in symmetric but weak
transit signals. For particularly fast planetary
outflows, a pre-transit dayside arm skewed
redward of line center is visible.
ii. Even a spatially and temporally steady inter-
mediate stellar wind cause periodic disruption
of the growing planetary outflow, leading to
large blue asymmetries preceding transit. A
26 Our simulations show that the fluid is still collisional through-
out the domain, due to Coulomb force's long range effect in the
mostly ionized outflow.
27
double transit-like feature, from the disrupted
dayside material, occurs at blue wavelengths.
iii. Strong stellar winds cause cometary-like plan-
etary outflows. Absorption is biased blue-
wards during transit, and the cometary tail-
like structure produces substantial blue ab-
sorption post-transit.
Absent from our discussion is radiation pressure from
Lyman-α, which at the extremes behaves either in an
optically thin or optically thick fashion. Recall that the
total fraction of incoming flux absorbed by neutrals is the
obscuration fraction, Oν = 1 − e−τν (spatially mapped
for our outflows in § 5.2). Thus, the total work done
by radiation pressure on the outflow is ∆Wν ∝ FνOν.
Optically thick radiation (τν (cid:29) 1 → Oν ≈ 1) absorbs
nearly all incoming flux at the edge of the outflow, and
behaves akin to a stellar wind. Optically thin radiation
(τν (cid:28) 1 → Oν ≈ τν) on the other hand, will only ab-
sorb a tiny fraction of the incoming flux, which occurs
throughout the entire outflow rather than at the edge.
In our simulations, the (unmodeled) force exerted by op-
tically thick photons would strongly dominate near the
planet, and would exceed that of optically thin photons
by at least a factor of two throughout the flow within our
domain. We note that optically thick radiation pressure
may behave similarly to stellar winds and be more appro-
priate to these systems. However, radiation pressure will
always act radially with respect to the star. In contrast,
the stellar wind has a ram-pressure headwind effect from
the planet's orbit and a thermal-pressure force normal to
the planetary-stellar wind contact discontinuity -- neither
of which are strictly radially outwards from the star.
In future work, one can use the machinery presented
here to analyze various stellar spectra, stellar time vari-
ability (particularly important for M-stars), planetary
and stellar magnetic fields, and the entrainment of heav-
ier elements. Discussed in § A.3, heat conduction should
be included to allow for more realistic stellar winds, i.e.,
coronal winds, and because it plays an important role
on the planetary nightside. As previously mentioned,
magnetic fields will play a dominant roll in shaping the
outflow if planetary fields exceed 1 G. As most of the
outflow is ionized, the planetary or stellar magnetic field
may prevent certain features in the large-scale outflow
from existing. Therefore, we emphasize again that our
results will reflect systems with weak magnetic fields, and
future works wishing to survey wider possibilities should
implement them.
This material is based upon work supported by the
National Science Foundation under Grant No. AST-
1411536. Further thanks to Jim Stone, James Owen,
and Jonathan Carroll-Nellenback for insightful numeri-
cal discussions. We would also like to acknowledge the
yt-project (Turk et al. 2011), which made possible the
majority of our visualizations.
REFERENCES
Barnes, J. R., Haswell, C. A., Staab, D., & Anglada-Escud´e, G.
2016, MNRAS, 462, 1012, doi: 10.1093/mnras/stw1713 [1]
Beckwith, K., & Stone, J. M. 2011, ApJS, 193, 6,
doi: 10.1088/0067-0049/193/1/6 [3.1.1]
McCann et al.
Ben-Jaffel, L. 2007, ApJL, 671, L61, doi: 10.1086/524706 [1, 5.4]
Ben-Jaffel, L., & Sona Hosseini, S. 2010, ApJ, 709, 1284,
doi: 10.1088/0004-637X/709/2/1284 [1]
Bisikalo, D., Kaygorodov, P., Ionov, D., et al. 2013, ApJ, 764, 19,
doi: 10.1088/0004-637X/764/1/19 [1]
Black, J. H. 1981, MNRAS, 197, 553,
doi: 10.1093/mnras/197.3.553 [2.1.1]
Bourrier, V., Lecavelier des Etangs, A., Ehrenreich, D., Tanaka,
Y. A., & Vidotto, A. A. 2016, A&A, 591, A121,
doi: 10.1051/0004-6361/201628362 [1, 5.4]
Bourrier, V., Lecavelier des Etangs, A., Dupuy, H., et al. 2013,
A&A, 551, A63, doi: 10.1051/0004-6361/201220533 [1, 5.4]
Carroll-Nellenback, J., Frank, A., Liu, B., et al. 2017, MNRAS,
466, 2458, doi: 10.1093/mnras/stw3307 [1, 2.3.4, 4.1, 14]
Chamberlain, J. W. 1961, ApJ, 133, 675, doi: 10.1086/147070
[A.1, A.1]
Cohen, O., Kashyap, V. L., Drake, J. J., et al. 2011, ApJ, 733, 67,
doi: 10.1088/0004-637X/733/1/67 [1]
Debrecht, A., Carroll-Nellenback, J., Frank, A., et al. in press [1,
4.1, 4.3, 4.7]
Draine, B. T. 2011, Physics of the Interstellar and Intergalactic
Medium [2.1.1]
Ehrenreich, D., Lecavelier Des Etangs, A., H´ebrard, G., et al.
2008, A&A, 483, 933, doi: 10.1051/0004-6361:200809460 [1]
Ehrenreich, D., Bourrier, V., Bonfils, X., et al. 2012, A&A, 547,
A18, doi: 10.1051/0004-6361/201219981 [1, 5.4]
Ehrenreich, D., Bourrier, V., Wheatley, P. J., et al. 2015, Nature,
522, 459, doi: 10.1038/nature14501 [1, 3, 5.4]
Fulton, B. J., Petigura, E. A., Howard, A. W., et al. 2017, AJ,
154, 109, doi: 10.3847/1538-3881/aa80eb [1]
Garc´ıa Munoz, A. 2007, Planet. Space Sci., 55, 1426,
doi: 10.1016/j.pss.2007.03.007 [1, 4.1]
Gombosi, T. I., van der Holst, B., Manchester, W. B., & Sokolov,
I. V. 2018, Living Reviews in Solar Physics, 15, 4,
doi: 10.1007/s41116-018-0014-4 [2.3.4]
Holmstrom, M., Ekenback, A., Selsis, F., et al. 2008, Nature, 451,
970, doi: 10.1038/nature06600 [1]
Jensen, A. G., Redfield, S., Endl, M., et al. 2012, ApJ, 751, 86,
doi: 10.1088/0004-637X/751/2/86 [1]
Krumholz, M. R., Stone, J. M., & Gardiner, T. A. 2007, ApJ,
671, 518, doi: 10.1086/522665 [3.1, 3.1.1]
Kulow, J. R., France, K., Linsky, J., & Loyd, R. O. P. 2014, ApJ,
786, 132, doi: 10.1088/0004-637X/786/2/132 [1]
Lamers, H. J. G. L. M., & Cassinelli, J. P. 1999, Introduction to
Stellar Winds (Cambridge University Press), 452 [2.2, 2.3.4,
A.3, A.3]
Lavie, B., Ehrenreich, D., Bourrier, V., et al. 2017, A&A, 605, L7,
doi: 10.1051/0004-6361/201731340 [1]
Lecavelier Des Etangs, A., Ehrenreich, D., Vidal-Madjar, A., et al.
2010, A&A, 514, A72, doi: 10.1051/0004-6361/200913347 [1]
Linsky, J. L., Yang, H., France, K., et al. 2010, ApJ, 717, 1291,
doi: 10.1088/0004-637X/717/2/1291 [1]
Lopez, E. D., & Fortney, J. J. 2013, ApJ, 776, 2,
doi: 10.1088/0004-637X/776/1/2 [1]
Loyd, R. O. P., Koskinen, T. T., France, K., Schneider, C., &
Redfield, S. 2017, ApJL, 834, L17,
doi: 10.3847/2041-8213/834/2/L17 [1]
Matsakos, T., Uribe, A., & Konigl, A. 2015, A&A, 578, A6,
doi: 10.1051/0004-6361/201425593 [1]
Miki´c, Z., Linker, J. A., Schnack, D. D., Lionello, R., & Tarditi,
A. 1999, Physics of Plasmas, 6, 2217, doi: 10.1063/1.873474
[2.3.4]
Murray, C. D., & Dermott, S. F. 1999, Solar system dynamics
[5.1]
Murray-Clay, R. A., Chiang, E. I., & Murray, N. 2009, ApJ, 693,
23, doi: 10.1088/0004-637X/693/1/23 [1, 2.1.1, 2.3.2, 2.4, 12,
3.2.2, 4.1, 4.7, 5.3, 5.4]
Oklopci´c, A., & Hirata, C. M. 2018, ApJL, 855, L11,
doi: 10.3847/2041-8213/aaada9 [1]
Osterbrock, D. E. 1989, Astrophysics of gaseous nebulae and
active galactic nuclei (University Science Books) [2.1.1, 2.1.1]
Owen, J. E., & Adams, F. C. 2014, MNRAS, 444, 3761,
doi: 10.1093/mnras/stu1684 [1]
Owen, J. E., & Wu, Y. 2013, ApJ, 775, 105,
doi: 10.1088/0004-637X/775/2/105 [1]
Parker, E. N. 1958, ApJ, 128, 664, doi: 10.1086/146579 [1, A.1]
28
McCann et al.
Pierrehumbert, R. T. 2010, Principles of Planetary Climate
Verner, D. A., Ferland, G. J., Korista, K. T., & Yakovlev, D. G.
[2.3.2]
1996, ApJ, 465, 487, doi: 10.1086/177435 [2.1.1]
Schneiter, E. M., Esquivel, A., Villarreal D'Angelo, C. S., et al.
Vidal-Madjar, A., Lecavelier des Etangs, A., D´esert, J.-M., et al.
2016, MNRAS, 457, 1666, doi: 10.1093/mnras/stw076 [1]
Seager, S., & Sasselov, D. D. 2000, ApJ, 537, 916,
doi: 10.1086/309088 [1]
2003, Nature, 422, 143, doi: 10.1038/nature01448 [1, 5.4]
-- . 2008, ApJL, 676, L57, doi: 10.1086/587036 [1, 24]
Vidal-Madjar, A., D´esert, J.-M., Lecavelier des Etangs, A., et al.
Spake, J. J., Sing, D. K., Evans, T. M., et al. 2018, Nature, 557,
2004, ApJL, 604, L69, doi: 10.1086/383347 [1, 5.4]
68, doi: 10.1038/s41586-018-0067-5 [1]
Stone, J. M., & Gardiner, T. A. 2010, ApJS, 189, 142,
doi: 10.1088/0067-0049/189/1/142 [3.1]
Stone, J. M., Gardiner, T. A., Teuben, P., Hawley, J. F., &
Simon, J. B. 2008, ApJS, 178, 137, doi: 10.1086/588755 [3.1]
Stone, J. M., & Proga, D. 2009, ApJ, 694, 205,
doi: 10.1088/0004-637X/694/1/205 [1, 2.3.4, 4.1, 5.2.3, 20]
Trammell, G. B., Li, Z.-Y., & Arras, P. 2014, ApJ, 788, 161,
doi: 10.1088/0004-637X/788/2/161 [1]
Tremblin, P., & Chiang, E. 2013, MNRAS, 428, 2565,
doi: 10.1093/mnras/sts212 [1, 2.3.4]
Vidal-Madjar, A., Huitson, C. M., Bourrier, V., et al. 2013, A&A,
560, A54, doi: 10.1051/0004-6361/201322234 [1]
Vidotto, A. A., & Bourrier, V. 2017, MNRAS, 470, 4026,
doi: 10.1093/mnras/stx1543 [3.3]
Vidotto, A. A., Lichtenegger, H., Fossati, L., et al. 2018, MNRAS,
481, 5296, doi: 10.1093/mnras/sty2130 [2.3.4, 3.2.2, 3.3, 5.4]
Watson, A. J., Donahue, T. M., & Walker, J. C. G. 1981, Icarus,
48, 150, doi: 10.1016/0019-1035(81)90101-9 [A.1]
Woods, T. N., Rottman, G. J., Bailey, S. M., Solomon, S. C., &
Worden, J. R. 1998, Sol. Phys., 177, 133,
doi: 10.1023/A:1004912310883 [3.3]
Tripathi, A., Kratter, K. M., Murray-Clay, R. A., & Krumholz,
Yelle, R. V. 2004, Icarus, 170, 167,
M. R. 2015, ApJ, 808, 173, doi: 10.1088/0004-637X/808/2/173
[1, 3.1, 3.3, 4.2, 4.7, C]
Turk, M. J., Smith, B. D., Oishi, J. S., et al. 2011, ApJS, 192, 9,
doi: 10.1088/0067-0049/192/1/9 [6]
APPENDIX
A. BERNOULLI CONSTANT
A.1. Overview of Bernoulli's Constant
The formalism of Bernoulli's constant is useful since it
gives analytic solutions to the atmospheric escape prob-
lem, potentially negating the need for numerical simula-
tions. One widely cited example of a quasi-analytic solu-
tion is Watson et al. (1981), who used conduction and a
delta function for ionization heating to simulate hydrody-
namic winds of our solar system's early terrestrial plan-
ets.27 One method of retrieving the Bernoulli's constant
is by integrating the energy equation along streamlines.
Equation (14) can be rewritten in steady state as
(cid:18) E + P
ρ
(cid:126)u · (cid:126)∇
(cid:19)
+ φ
=
G − L
ρ
.
Take for now the following as the definition for ∆q
(cid:126)u · (cid:126)∇ (∆q) ≡ G − L
ρ
.
Then by substituting Eq. (A2) into Eq. (A1) and inte-
grating along streamlines one retrieves the constant
b =
1
2
u2 + h + φ − ∆q.
(A3)
Here we have made use of the specific enthalpy, h =
+ P/ρ, where is the specific internal energy. Equa-
tion (A3) is the specific Bernoulli constant, which is phys-
ically motivated by conservation of energy yet is not the
total specific energy. Rather b is the total energy minus
the total heat added to the fluid. When external heat is
added to the fluid, it becomes either kinetic, internal or
potential energy, so for b to remain constant the added
energy must be subtracted off the total energy.
At first glance it may seem challenging to invert
Eq. (A2); however, recall that the convective derivative,
27 They used numerics to iteratively solve the steady state sys-
tem of equations (information contained in the Bernoulli constant
plus further assumptions), as opposed to integrating the fluid equa-
tions forward to a steady state.
(A1)
(A2)
doi: 10.1016/j.icarus.2004.02.008 [1]
or directional derivative, ∇u = (cid:126)u· (cid:126)∇, differentiates a field
along streamlines of the flow. This can be thought of as
a 1-D problem, e.g., by using the Frenet-Serret frame, so
that ∆q can be solved as
∆q =
C(s)
G − L
ρus
ds.
(A4)
(cid:90)
Here C(s) is the streamline parameterized by s. Note
that by definition in these coordinates us = (cid:126)u.
This prescription for Bernoulli's constant can be ex-
panded upon to include a wide range of physics. Typ-
ically one describes the means of energy gain as either
heat gained by or work done on the fluid. While conser-
vative forces should be included in φ, non-conservative
forces, i.e., ones with path-dependent work, require a
path-dependent term, ∆W , measuring the accumulated
work done along the path. We do not consider such forces
in this work, but note that radiation pressure would be
such a term.
Another common example of heat flux through a fluid
is conduction. One can quickly see that if conduction is
redistributing heat that
(cid:90)
(cid:17)
(cid:126)∇ ·(cid:16)
κ (cid:126)∇T
ρus
∆q =
C(s)
ds.
(A5)
The classic example of heat conduction in atmospheric
escape,
is Parker's spherically symmetric stellar wind
model (1958), also see Chamberlain (1961). Recall that
for spherical symmetry the constant mass-loss rate can
be expressed as M = ρur2. Then Eq. (A5) can be solved
so that
b =
1
2
u2 + h + φ − κ
ρu
∂T
∂r
.
(A6)
This is equivalent to Eq. 12 of Chamberlain (1961). In
this special case, an integral accounting for the accumu-
lated heat is not needed and only local quantities matter.
Since in a spherically symmetric system heat conduction
can act only along streamlines, it cannot alter the total
energy along a streamline except at the boundaries. This
however is not generally true when the assumption of
29
spherical symmetry is relaxed, as heat flux from conduc-
tion may now flow perpendicular to streamlines -- making
∆q path-dependent.
A.2. Reversible flows
A perhaps more intuitive derivation of Eq. (A4) for
∆q results from considering how the energy equation is
constructed from Euler's equation, Eq. (13), and the fun-
damental thermodynamic relationship for enthalpy. We
use the powerful simplification that the flows are well ap-
proximated as reversible, so that a flow experiencing only
reversible pressure-volume work has a specific enthalpy
(cid:18) ρ
(cid:19)Γ
ρ0
deposition in the stellar wind. This assumption allows
us the relationship
McCann et al.
P = P0
.
(A11)
In order to retrieve Bernoulli's constant from the Eu-
ler equation, as discussed in § A.2, we can reduce the
pressure force per unit mass to a scalar gradient via the
polytropic relationship, for Γ (cid:54)= 1
(cid:18) P
(cid:19)
ρ
(cid:126)∇P
ρ
=
Γ
Γ − 1
(cid:126)∇
.
(A12)
dh = ¯dq +
dP
ρ
.
(A7)
So then
Here ¯dq is the differential heat flow into the system, e.g.,
a fluid element. Recall that ¯dq is a form of energy transfer
and must be represented by an inexact differential, i.e.,
no path-independent integrated quantity, say q, is well
defined. However, if a path, C, parameterized by s, is
specified we can define the accumulated heat along said
path
bpoly =
1
2
u2 +
Γ
Γ − 1
P
ρ
+ φ.
For an ideal gas
h =
γ
γ − 1
P
ρ
,
(cid:90)
C(s)
∆q(s) ≡
¯dq.
(A8)
and
(A13)
(A14)
(A15)
For a fluid element the obvious path would be its
streamline. Consider a fluid in a steady state, ∂t (∆q) =
0, then it is clear that the infinitesimal heat added at s,
¯dq(s), is the convective derivative of ∆q, i.e.,
(cid:104)
(cid:105)(cid:12)(cid:12)(cid:12)s
¯dq(s) =
(cid:126)u · (cid:126)∇ (∆q)
.
(A9)
This is also apparent from Eq. (A8). Therefore consid-
ering Eq. (A7) as material derivatives
(cid:126)u · (cid:126)∇P
ρ
= (cid:126)u · (cid:126)∇h − (cid:126)u · (cid:126)∇(∆q).
(A10)
The term on the left-hand side of Eq. (A10) appears when
you project Euler's equation into streamlines, a close
analog to the energy equation.
Integrating that equa-
tion along the streamline likewise retrieves the Bernoulli
constant for a compressible flow with external heating,
Eq. (A3), which we have now shown is completely gen-
eral as long as P dV work done by the flow is reversible,
i.e., the only entropy production comes from heating.
We have thus arrived at a more natural definition for
∆q in Eq. (A8). Typically ¯dq is understood in terms of
a heating rate, which we would integrate with respect to
time to get the total heat added. Referencing Eq. (A4),
we note that ds/us is the time interval the fluid element
spends near a given point and the rest of the integrand
is the specific heating rate.
A.3. Polytropic flows
bpoly = bisen +
γ − Γ
(γ − 1)(Γ − 1)
P
ρ
,
where bisen is the isentropic Bernoulli constant (Eq. (A3)
with ∆q = 0). Note that bisen is typically thought of as
the total energy.
Therefore, when Γ = γ, there is no bound atmosphere
(bisen < 0) that can become unbound (bisen ≥ 0), and
"adiabatic escape" is a misnomer. A correlated point is
that there exist energy (or momentum transfer) require-
ments at the sonic point for the outflow to go transonic.
In other words, it is not only enough that energy be de-
posited in the fluid, but there must also be non-zero en-
ergy deposition below some threshold at the sonic point
for the solution to be transonic (too large of an energy
input at the sonic point reduces the outward pressure
force, lessening the accelerating of the outflow). The
justification of this constraint requires careful analysis
of the momentum equation and is given in Lamers and
Cassinelli (1999) § 4.1.4 for spherical winds.28
As discussed in § A.1, the Bernoulli constant can be
used to analytically solve atmospheric escape. How-
ever, the application has its limitations, namely the con-
stant knows nothing of the ionization structure that di-
rectly corresponds to ∆q. This makes outflows that are
launched by ionization heating, or for which you wish
to make synthetic observations, not accessible by the
Bernoulli constant alone. However, for our stellar wind
model we do not care about the ionization structure, and
use the Bernoulli constant to derive its solution. We
model the stellar wind as a spherically-symmetric, isen-
tropic wind (polytropic outflow with Γ = γ). To solve we
The Bernoulli constant provides a useful frame for un-
derstanding the setup of our stellar wind boundary con-
dition, which we model as a polytropic outflow.
It is
often common to neglect detailed treatment of energy
28 We note that Eq. (4.31) in the reference is missing a minus
sign and should read −∂r log(qr−1/2) > (5/8)∂rM 2 > 0 when
evaluated at rc, to be in agreement with both Eq. (4.29) and the
discussion following.
30
begin by writing the steady state mass continuity equa-
tion as
ρur2 = ρ0u0r2
0.
(A16)
McCann et al.
This,
combined with the polytropic relationship,
Eq. (A11), allows us to write the sound speed in terms
of the velocity and radius
(cid:115)
(cid:19)γ−1
(cid:18) u0r2
0
ur2
cs(r) =
γP0
ρ0
.
(A17)
Combined with Eq. (A13) we have an equation for u in
terms of only constants and r
(cid:19)γ−1
(cid:18) u0r2
0
ur2
1
2
u2 +
γ
γ − 1
P0
ρ0
+ (φ − bpoly) = 0.
(A18)
Thus given a mass-loss rate,
M = ρur2, and the tem-
perature of the outflow at r0, we can analytically solve
for the velocity structure of the outflow. This is precisely
what we do for our stellar wind, by using Brent's method
to solve for u at all locations within the domain. We then
take the velocity and use Eqs. (A16) and (A17) to solve
for the density and temperature structure of the stellar
wind.
Note that for future work one should consider instead
using what is sometimes called a coronal wind (Lamers
& Cassinelli 1999, Ch. 5). We have already derived the
solution for the coronal wind in Eq. (A6). Notice that
for coronal winds b = bisen − (κ∂rT )/(ρu). Thus, if one
has large heat fluxes at the wind base (large −κ∂rT
and/or small ρu) it is possible to have bound atmo-
sphere (bisen < 0) near the wind base, but be unbound
(bisen ≥ 0) further out where heat fluxes have distributed
the heat (−κ∂rT ≈ 0). Importantly, this can be accom-
plished in simulations with γ = 5/3, meaning that with
conduction, transonic stellar winds can be included in
simulations that include adiabatically cooled planetary
outflows.
B. AMBIENT MEDIUM SETUP AND OPTIMIZATION
Illustration of radial scales for the nested isentropic
Figure 21.
atmospheres that comprise our initial conditions for simulations
that lack a stellar wind. Blue denotes our reference height for each
atmosphere and black is the bounding surface of that atmosphere
where the values go to zero. Not to scale.
As mentioned in § 2.3.2, an isentropic bound hydro-
static atmosphere has a bounding surface at which the
density and pressure go to zero. To avoid the difficulty
of handling this numerically, our atmospheres are cut
off at a numerical edge such that Re < Rz. It is desir-
able to chose Re well away from the τ = 1 surface for
photoionization, This prevents unnecessary evolution of
τ = 1, although the final solution should be insensitive
to any such choices. More importantly, the closer Re is
to Rz, the lower the pressure and density of the pressure-
matched ambient medium can be. We find that setting
the Re = Rz − δx/2 produces numerical stability, where
δx is the size of a computational cell at the locally high-
est level of static mesh refinement. This choice is suc-
cessful because as part of the hydrodynamic update we
must extrapolate the state variables from cell centers to
cell edges. Truncating at Rz − δx/2 ensures that a well-
defined solution exists at the cell edge to which we are
extrapolating. Note that our stellar wind is also techni-
cally an isentropic atmosphere; however, it is unbounded
so that it does not have a zero radius, Rz. Therefore, we
only consider a reference radius, r(cid:63),0, at which to pre-
scribe boundary conditions.
B.1. Setup
B.2. Optimization
When no stellar wind is present, our ambient medium
takes the form of two nested isentropic atmospheres,
which prevents the background gas from collapsing due
to the planet's gravity. Our simulations thus contain up
to three nested isentropic atmospheres, pressure-matched
at the boundaries where they meet. We enumerate our
isentropic atmospheres as follows: "0" corresponds to
the planetary atmosphere, "1" to the primary ambi-
ent medium and "2" to the secondary ambient medium.
Each isentropic atmosphere has a reference radius R0,
R1, or R2, a zero radius Rz, where the atmosphere for-
mally reaches zero density, and a numerical edge, Re,
where the atmosphere is truncated in our simulation.
For example, Re,0 refers to the truncation radius for the
planetary atmosphere. As the atmospheres are nested
and pressure matched, we take the edge of an inner at-
mosphere as the reference of the next atmosphere, e.g.,
R1 = Re,0. These radial scales are illustrated in Fig-
ure 21.
The numerical difficulties in simulating our ambient
medium are twofold. First, for numerical stability, areas
with large gradients should be well resolved. To effi-
ciently achieve this we need to avoid regions of the isen-
tropic ambient medium that have small scale heights,
e.g., near their zero radius. The second difficulty comes
from pressure matching a low- and high-density region,
since the speed of sound in the low density region will
be larger. Lowering the ambient density then in turn,
leads to exceedingly expensive computations as we must
satisfy the Courant condition.
When trying to optimize our simulations, the limita-
tions placed by these two constraints turn out to be at
odds with one another. Since we must prioritize accu-
racy over efficiency, we foremost avoid large gradients in
the flow, and we will treat the sound speed as a sec-
ondary optimization to consider only if we can ensure
small gradients. Why these constraints are at odds with
one another is intuitively understood by considering the
31
atmosphere's temperature. A hotter atmosphere has a
larger sound speed, which by the Courant condition re-
quires higher temporal resolution. Conversely a colder
atmosphere has smaller scale heights that requires higher
spatial resolution to resolve. As the density is related to
the temperature by the equation of state, we will now ex-
amine the bounds these constraints place on our ambient
density.
As already stated, the first difficulty (large gradients) is
averted by avoiding the edge of the ambient atmosphere.
From Eq. (24) we know Rz,1 occurs where φ(Rz,1) =
h1 + φ1. To avoid a "snowplow" phase as our planetary
wind launches, we wish to reduce the reference density of
the ambient medium, ρ1, relative to that of the edge of
the planetary atmosphere, ρ(Re,0). Then, as the ambient
medium and planetary atmosphere are pressure matched
(P1 = P (Re,0)), h1 > h(Re,0) and thus Rz,1 > Rz,0 (see
Footnote 10 for detailed proof). However, rather than
setting ρ1 and solving for Rz,1, we instead solve for ρ1
after choosing Rz,1. Operationally we chose the Rz,1 to
either be outside of our domain or at some special radius,
as discussed below.
To bound ρ1, consider a potential such that φ(r) ≥ φ1
for r > R1. This happens for a point mass or, in the
full rotating reference frame, inside the Roche lobe. To
ensure that the density does not go to zero inside Re,1,
we require φ(Re,1) > φ(Rz,1), or
ρ1 ≤
γP1
(γ − 1) (φ(Re,1) − φ1)
.
(B1)
For the lower bound, it turns out that inside our isen-
tropic planetary atmosphere the largest sound speed is
found in the inner masked region. Therefore, we seek an
ambient medium that at most has this speed of sound
to add no further computational expense. Using the
Courant condition to keep the adiabatic sound speed be-
low that of the masked region, the ambient gas density
should satisfy
ρ1 ≥ ρmask
P1
Pmask
.
(B2)
Here the "mask" subscript is the value inside Rmask.
Combining Eqs. (B1) and (B2) provides upper and lower
constraints on our ambient density
ρmask
P1
Pmask
≤ ρ1 ≤
γP1
(γ − 1) (φ(Rz,1) − φ1)
.
(B3)
Ideally, one should chose the lowest density possible
to avoid causing the wind to enter a "snowplow" phase
when expanding. Yet, there is no guarantee that these
constraints are consistent, and as mentioned we must fa-
vor satisfying Eq. (B1). If these constraints turn out to
be inconsistent then there is additional computational
expense in our ambient medium over the planetary at-
mosphere. However, since we are using structured static
mesh refinement we can further exploit the larger cell
sizes in the lower resolution regions to perhaps make in-
consistent bounds consistent, or at the very least "less"
inconsistent.
Let the cell size ratio between the coarsest and finest
resolution be f . By the Courant condition on the coarser
32
McCann et al.
mesh we can have a factor of f larger sound speed com-
pared to the fine resolution mesh, which translates into
a factor of f 2 lower density. Therefore, if we extend our
first ambient medium layer out only to the coarse refine-
√
ment, e.g., if it is given by a cube of side 2Rrefine then
3Rrefine, we can use a factor of f 2 smaller density here
while satisfying the Courant conditions with the same
time step. Then from there a second ambient medium
layer would extend out to infinity, Rz,2 → ∞, so that
φ(Rz,2) = 0. Then the constraints on our secondary am-
bient medium layer become
(cid:18)
(cid:19)
1
f 2
ρmask
P2
Pmask
≤ ρ2 ≤
γP2
(γ − 1) (−φ2)
.
(B4)
√
Here the subscript "2" denotes the second ambient
layer's reference point, which is at
3Rrefine = Rz,1.
There is still no guarantee that these conditions are con-
sistent. However, by exploiting the refinement factor
and using multiple ambient layers one can optimize their
initial setup relative to a single ambient layer. This is
only appropriate when lower resolution does not lead to
numerical instability or loss of numerical convergence.
We therefore need the scale height of our wind to be
larger than our isentropic hydrostatic atmosphere, which
is what we find.
C. SCALE HEIGHTS WITHIN THE ATMOSPHERE
We will now review four distinct definitions of a scale
height and explore their uses. For simplicity we will work
with a generic variable X in a spherically symmetric at-
mosphere. Often the first scale height one encounters,
and perhaps the best defined,
is the isothermal scale
height, Hiso = kBT /(µg). It is derived from the plane-
parallel, isothermal atmosphere, for which the thermo-
dynamic variables are of the form X(z) = X0 exp(−(z −
z0)/Hiso). Thus, for each Hiso away from z0, the ther-
modynamic variables have an e -- folding in value.
The second way in which we will define a scale height
is an often-used order of magnitude definition that can
be thought of as the natural extension of isothermal scale
height
(cid:18)
(cid:19)−1
H(cid:48) =
− d log(X)
.
dr
(C1)
Solving this differential equation for constant H(cid:48) re-
trieves the isothermal atmosphere. However, this def-
inition is often used even when the atmosphere is not
an exponential, i.e., H(cid:48) is not constant. For example, we
have used this definition for our length scale in our Knud-
sen number calculation, (L = (∇ log P )−1). That is be-
cause it roughly encapsulates the scale on which the vari-
ables change the order of themselves (H(cid:48) ∼ (X/∆X)∆r).
When one integrates Eq. (C1), it counts the number of
e -- foldings between two locations,
(cid:90) b
(cid:18) X(b)
(cid:19)
Ne(a, b) = −
H(cid:48)−1dr = log
(C2)
but what does H(cid:48)(r) itself mean in an atmosphere that
is not an isothermal plane-parallel atmosphere? One will
find that in general it is not the distance from r at which
X(a)
a
,
variables have e -- folded. Instead it describes the e -- folding
length in the analogous isothermal plane -- parallel atmo-
sphere that has the same conditions at r as our atmo-
sphere. If we want the scale height, H(r), over which the
atmosphere folds by a factor of s, we need to solve
X(r + H(r)) = sX(r).
(C3)
This is usually what we mean when we talk about a
scale height in our isentropic atmospheres, since it is
useful to determine the actual folding within a numer-
ical cell to quantify a condition for numerical stability.
Even for a spherically symmetric isothermal atmosphere
the form is not simple; however, both the isothermal
and polytropic spherically symmetric cases have closed
forms. For polytropic atmospheres all thermodynamic
variables are a power, n, of an underlaying function f (r),
i.e., X(r) = X0f (r)n, e.g., n = 1/(γ − 1) for density,
n = γ/(γ − 1) for pressure, and n = 1 for temperature.
Then for the corresponding variable to the index n, the
scale height H(r) has a s-folding after
Hn(r, s) =
Rz − r
r + Rz(s−n−1 − 1)−1
r,
(C4)
where Rz is the edge of your atmosphere, which for a hy-
drostatic isentropic atmosphere in a point mass potential
is given by
(cid:32)
(cid:33)−1
Rz =
−
1
R0
γc2
s,0
(γ − 1)GMp
.
(C5)
The last way which we define a scale height only applies
for atmospheres that tend to zero in the limit of infinity.
It is defined as(cid:90) ∞
r
X(r(cid:48))dr(cid:48) = X(r)(cid:101)H(r).
(C6)
Clearly this scale height is undefined in atmospheres that
do not vanish at infinity, such as the spherically symmet-
ric isothermal atmospheres.29 For a hydrostatic poly-
tropic atmosphere, this scale height has a closed form in
McCann et al.
(cid:18) Rz
(cid:19)−3/2(cid:18)
terms of incomplete beta functions
(cid:19)(cid:19)
(cid:101)H(r) = Rz
This last scale height, (cid:101)H, is useful since its definition
, 1 − n, 1 + n
(C7)
(cid:18) r
− 3π
2
− β
− 1
Rz
.
parallels that of optical depth
σ nHI(r(cid:48)) dr(cid:48) = σ nHI(r) (cid:101)H(r).
(C8)
τ (r) =
r
(cid:90) ∞
r
Therefore, rather than doing a numerical integral to cal-
culate the optical depth in our isentropic atmospheres,
we only need to calculate the incomplete beta function.
We use this calculation when picking an initial number
density at Rp (Eq. (30)). This is can be justified by
considering where τ = 1 occurs in steady state.
From Eq. (C8), one can always pick a nHI(Rp) such
that τ = 1 at Rp in steady state. However, the issue
is knowing a priori the steady-state (cid:101)H of the outflow.
For the energy-limited regime, where the outflow is not
strongly ionized and neutrals are replenished by advec-
tion rather than recombination, we find that the opti-
cal depth one surface does not change substantially from
its initial location. We calculate the number density re-
quired to place τ = 1 at Rp using Eq. (C7) and Eq. (C8).
However, in the radiation-limited regime the assump-
tion that the ionization front does not move breaks
down. We therefore no longer suggest using Eq. (C8),
and instead appeal to the fact that τ = 1 will occur
where recombination balances ionization ((1−X)σHIF =
αBnHX 2).30 Solving for nH gives
σHI F0
1 − X
X 2
αB
nH =
(C9)
For the ionization fraction a small value, perhaps X ∼
0.1 is appropriate, as in the radiation-limited regime the
ionization structure will be extremely sharp. The ideal
value of X merits future investigation.
.
Note that, we do not provide a first-order principles
method of knowing a priori if you are in the energy-
limited or radiation-limited regime, but it suffices to
say that from previous studies we knew our parameters
would firmly be in the energy-limited regime (Tripathi
et al. 2015).
three scale heights: H(cid:48), H, and (cid:101)H reduce to Hiso for
We conclude by noting, perhaps unsurprisingly, all
the plane-parallel isothermal atmospheres.
29 Note that a plane-parallel isothermal atmosphere may extend
infinity far out, but it tends to zero in the limit, so it is well defined
there.
30 This is not true in the energy-limited as advection primarily
balances ionization.
33
|
1712.05154 | 2 | 1712 | 2017-12-15T10:40:39 | On the impact origin of Phobos and Deimos III: resulting composition from different impactors | [
"astro-ph.EP"
] | The origin of Phobos and Deimos in a giant impact generated disk is gaining larger attention. Although this scenario has been the subject of many studies, an evaluation of the chemical composition of the Mars' moons in this framework is missing. The chemical composition of Phobos and Deimos is unconstrained. The large uncertainty about the origin of the mid-infrared features, the lack of absorption bands in the visible and near-infrared spectra, and the effects of secondary processes on the moons' surface make the determination of their composition very difficult from remote sensing data. Simulations suggest a formation of a disk made of gas and melt with their composition linked to the nature of the impactor and Mars. Using thermodynamic equilibrium we investigate the composition of dust (condensates from gas) and solids (from a cooling melt) that result from different types of Mars impactors (Mars-, CI-, CV-, EH-, comet-like). Our calculations show a wide range of possible chemical compositions and noticeable differences between dust and solids depending on the considered impactors. Assuming Phobos and Deimos as result of the accretion and mixing of dust and solids, we find that the derived assemblage (dust rich in metallic-iron, sulphides and/or carbon, and quenched solids rich in silicates) can be compatible with the observations. The JAXA's MMX (Martian Moons eXploration) mission will investigate the physical and chemical properties of the Maroons, especially sampling from Phobos, before returning to Earth. Our results could be then used to disentangle the origin and chemical composition of the pristine body that hit Mars and suggest guidelines for helping in the analysis of the returned samples. | astro-ph.EP | astro-ph | Draft version December 18, 2017
Typeset using LATEX default style in AASTeX61
ON THE IMPACT ORIGIN OF PHOBOS AND DEIMOS III:
RESULTING COMPOSITION FROM DIFFERENT IMPACTORS
Francesco C. Pignatale,1 S´ebastien Charnoz,1, 2 Pascal Rosenblatt,3, 4 Ryuki Hyodo,5 Tomoki Nakamura,6 and
Hidenori Genda5
1Institut de Physique du Globe de Paris (IPGP), 1 rue Jussieu, 75005, Paris, France
2Institut de Physique du Globe/Universite Paris Diderot/CEA/CNRS, 75005 Paris, France
3Royal Observatory of Belgium, Avenue circulaire 3, B-1180 Uccle, Belgium
4now at ACRI-ST, 260 route du pin-montard- BP 234- F-06904 Sophia-Antipolis Cedex, France
5Earth-Life Science Institute/Tokyo Institute of Technology, 152-8550 Tokyo, Japan
6Tohoku University, 980-8578 Miyagi, Japan
(Accepted for publication in ApJ)
ABSTRACT
The origin of Phobos and Deimos in a giant impact generated disk is gaining larger attention. Although this scenario
has been the subject of many studies, an evaluation of the chemical composition of the Mars' moons in this framework
is missing. The chemical composition of Phobos and Deimos is unconstrained. The large uncertainness about the
origin of the mid-infrared features, the lack of absorption bands in the visible and near-infrared spectra, and the
effects of secondary processes on the moons' surface make the determination of their composition very difficult from
remote sensing data. Simulations suggest a formation of a disk made of gas and melt with their composition linked
to the nature of the impactor and Mars. Using thermodynamic equilibrium we investigate the composition of dust
(condensates from gas) and solids (from a cooling melt) that result from different types of Mars impactors (Mars-,
CI-, CV-, EH-, comet-like). Our calculations show a wide range of possible chemical compositions and noticeable
differences between dust and solids depending on the considered impactors. Assuming Phobos and Deimos as result of
the accretion and mixing of dust and solids, we find that the derived assemblage (dust rich in metallic-iron, sulphides
and/or carbon, and quenched solids rich in silicates) can be compatible with the observations. The JAXA's MMX
(Martian Moons exploration) mission will investigate the physical and chemical properties of the Maroons, especially
sampling from Phobos, before returning to Earth. Our results could be then used to disentangle the origin and
chemical composition of the pristine body that hit Mars and suggest guidelines for helping in the analysis of the
returned samples.
Keywords: planets and satellites: composition, planets and satellites: formation, planets and satellites:
individual (Phobos, Deimos)
7
1
0
2
c
e
D
5
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
4
5
1
5
0
.
2
1
7
1
:
v
i
X
r
a
Corresponding author: Francesco C. Pignatale
[email protected]
2
Pignatale et al.
1. INTRODUCTION
The history of formation of the Mars' moons Phobos and Deimos is still an open question. It has been the subject of
several studies which point to a capture origin, in-situ or impact generated formation (Rosenblatt 2011; Citron et al.
2015; Rosenblatt et al. 2016, and references therein). Accretion within an impact-generated disk scenario (Craddock
2011; Rosenblatt et al. 2016) is gaining more support as it can explains several properties of the Mars' moons such as
the mass and the orbital parameters (Rosenblatt et al. 2016; Hesselbrock & Minton 2017; Hyodo et al. 2017a,b).
Phobos has a very peculiar infrared spectra. Although mid-infrared (MIDIR) show different features, the visibile
(VIS) and near-infrared (NIR) spectra are characterized by a lack of absorption features (Murchie 1999; Giuranna et al.
2011; Rosenblatt 2011; Murchie et al. 2015). Murchie (1999) isolated two main regions named "red" and "blue" on
the Phobos' surface that have different spectral characteristics which can be best matched by D- and T-type asteroids
respectively (Murchie et al. 1991; Murchie 1999; Rivkin et al. 2002). Giuranna et al. (2011) presented a detailed
investigation on the possible chemistry of Phobos' surface. They found that the "blue" region can be fitted with a
phyllosilicates-rich material, while the "red" region has a best fit when tectosilicates, mainly feldspar, are included in
the model. Moreover they found that no class of chondritic material can match the observed spectra. Nevertheless,
they pointed out that different more complex mixtures of dust could be able to reproduce the observed trends.
The featureless VIS-NIR spectra are often associated with a strong space weathering (Murchie 1999; Rosenblatt
2011). However, Giuranna et al. (2011) following the spectroscopical studies of Singer (1981), Salisbury & Walter
(1989),Cloutis et al. (1990a), Cloutis et al. (1990b), Cloutis et al. (1990c), Cloutis et al. (1990d), Burns (1993), Klima
et al. (2007) list a series of possible mechanisms that can reduce the strength of the spectra and match the observation:
i) as the 1-2µm feature arise from iron-bearing material such as pyroxene and olivine, the absence of those compounds
may reduce the spectra; ii) a mixture of opaque material such as metallic iron, iron oxides and amorphous carbon
mixed with olivine and pyroxene can reduce dramatically the VIS/NIR bands; iii) solids which results from quenching
from the liquids state may have their reflectance properties reduced as they lack of perfect crystalline structure; iv)
the reflectance of fine-grain materials decreases as the size of the grains decreases.
Hyodo et al. (2017a) presented detailed Smoothed Particle Hydrodynamics (SPH) simulations in which they deter-
mined the dynamical, physical and thermodynamical properties of an impact-generated disk. They found that the
material that populate the disk is initially a mixture of gas (∼ 5%) and melts (∼ 95%). These information together
with the Martian composition and hypothesis on the impactors, can be used for modelling the building blocks of
Phobos and Deimos.
In this work we present a study of the bulk composition of the Mars's moons following the giant-impact scenario.
Our aim is to provide more clues on the origin of the moons, their chemical composition, infrared spectra, and the
nature of the impactor itself.
Furthermore, the JAXA's MMX1 mission plans to observe Phobos and Deimos in detail, and return samples (at
least 10g) from the surface of Phobos. Our results could be then used as guidelines for helping in their analysis and
interpretation.
Starting from different initial compositions of the impactor (from mars-like to chondritic-like), we compute thermo-
dynamic equilibrium (DeHoff 1993) to solve for stable phases that may condense from the gas in the impact-generated
disk. Additionally, we compute the composition of the cooling melt to investigate how it will eventually differs from
condensates. The resulting condensates and solidified melt are then taken as proxies for the building block of Phobos
and Deimos and further discussion are made.
In this work we will mainly focus on Phobos, as more observation are available and as it will be the main sampling
target of the JAXAs MMX mission. Nevertheless, the formation of Deimos follows the same proposed scenario.
The paper is structured as follow: in section 2 we describe the techniques and the model we use in our calculations.
In section 3 we present our results that will be discussed in section 4. Conclusions are summarized in section 5.
Hyodo et al. (2017a) calculated that the temperature in the Mars' moons forming region of the disk reaches T ∼
2000 K just after the impact. The value of P ∼ 10−4 bar is chosen as our fiducial pressure as it is, for the given
temperature, the average saturation pressure for several mixtures calculated in Visscher & Fegley (2013) and the
average pressure in the disk profile in Ronnet et al. (2016) and Hyodo et al. (2017a) where gas and melt coexist.
2. MODEL AND METHODS
1 http://mmx.isas.jaxa.jp/en/index.html
AASTEX Dust composition from impactors
3
Figure 1. This cartoon describes the considered scenario. After the impact, part of Mars material will be ejected out at high
temperature and will vaporize into gas as well as part of the impactor. The gas mixture will then condense into dust. On the
other hand, the not vaporized material from Mars and the impactor will form a melt and then solidify. Phobos and Deimos will
be the result of the accretion of these two components. The yellow region represents the part of disk within the Roche limit
(Hyodo et al. 2017a).
Under these conditions the material in the disk that comes from Mars and from the impactor will result in a mixture
composed of gas and melt (Hyodo et al. 2017a).
Hyodo et al. (2017a) showed that the building blocks of Phobos and Deimos would be composed of a mixture of
about half-martian material and half-impactor material. We, thus, assume that the gas is made of a well mixed
two-components: the gas that is released by heating up Mars-material plus the gas that is released by heating up
impactor-material. We then assume that the melt is a mixture of the not vaporized material from the two bodies
(Hyodo et al. 2017a). Figure 1 shows a cartoon of the proposed model.
As the disk cools down, the gas will eventually re-condense and the melt will solidify. In this work, we define, for
ease of understanding, dust as the condensates from the gas phase and solids as the material that result from the
solidification of the melt.
In order to determine the composition of the dust that will condense from the gas phase we assume thermodynamic
equilibrium (DeHoff 1993): at constant temperature and pressure, the stability of a system is determined by its
Gibbs free energy, and, in fact, by the composition which minimizes the potential of the system. Although it is an
approximation, thermodynamic equilibrium is a powerful tool to understand the evolution of the chemical composition
of complex systems. This technique has been extensively used in the study of the chemistry of gas and dust in several
astrophysical environments:
from the Solar Nebula, meteorites and protoplanetary disks (Larimer 1979; Yoneda &
Grossman 1995; Lodders 2003; Ebel 2006; Pignatale et al. 2016) to stars dusty envelopes (Gail & Sedlmayr 1999;
Lodders & Fegley 1997; Ebel & Grossman 2001) and exoplanets composition (Bond et al. 2010).
4
Pignatale et al.
To compute the thermodynamic equilibrium we use the HSC software package (version 8) (Roine 2002), which
includes the Gibbs free energy minimisation method of White et al. (1958). Thermodynamic data for each compound
are taken from the database provided by HSC (Roine 2002, and references therein). HSC has been widely used
in material science and it has been already tested in astrophysics showing very good reliability in predicting the
composition of different systems (Pasek et al. 2005; Pignatale et al. 2011; Bond et al. 2010; Madhusudhan et al. 2012).
To calculate the composition of the solids from the cooling melts we use the normative mineralogy (CIPW-norm)
(Cross et al. 1902) and the work of Ronnet et al. (2016) as benchmark. CIPW-norm is one of the most used technique
to determine, in a first approximation, the equilibrium composition of a multicomponent melt (Cross et al. 1902).
(Hyodo et al. 2017a) showed that the melt phase of Mars and the impactor will likely never completely equilibrate
between each other. Mars-only and impactor-only melt with different degrees of equilibration in between are indeed
expected. Nevertheless, calculating the resulting compositions of a equilibrated melt represents a first interesting add-
on to investigate the differences that condensation and solidification would bring to the final Phobos bulk composition.
Moreover, our model suggests that the MMX may be confronted with two distinguishable family of material, the dust
and the solids. As a consequence, this investigation can bring further information and clues that can be used in the
MMX samples analysis.
During the planet formation, Mars and the other inner rocky planets experienced impacts with other bodies. The
impact histories strongly depend on the timing and location of the planets (Brasser et al. 2017; Raymond & Izidoro
2017; Brasser & Mojzsis 2017; Bottke & Andrews-Hanna 2017). The nature of the impactors is unconstrained as the
dynamical interactions of Jupiter and Saturn with the surrounding minor bodies may have scattered and delivered
in the inner Solar System material of different nature and of chondritic origin (Raymond & Izidoro 2017; Brasser &
Mojzsis 2017; Bottke & Andrews-Hanna 2017). Our aim is to determine the changes that different impactors would
bring in the chemical composition of Phobos, and if these differences can be traceable. In order to keep our selection
as chemically heterogeneous as possible we, thus, consider the following types of impactor: Marstype, CVtype, EHtype,
CItype, comettype. As a proxy of Mars composition we take the Bulk-Silicates-Mars (BSM) from Visscher & Fegley
(2013). Compositions for the EH, CV and CI chondrites are taken from Wasson & Kallemeyn (1988). Elemental
composition for the comet is taken from Huebner & Boice (1997). Table 1 shows the elemental distribution for all
considered impactors. In order to help to understand the differences between the impactors we also report several
elemental ratios such as the Mg/Si, Fe/O, C/O ratios. These ratios play an important role in determining the resulting
chemical composition of a mixture. This will be discussed in section 4.
We also report values from the Sun photosphere2 as reference (Asplund et al. 2009). Note the H/O ratio of the solar
nebula (Sun) and the abundances of other elements relatives to O and C. Looking at table 1, we can already notice that
we will deal with wide different environments. Moreover, the relative abundances between elements clearly indicate
that our systems will return chemical distributions that are far from that one predicted for a solar composition.
The 13 considered elements in Table 1 can form ∼6800 possible compounds, including complex organics, gas and
solids (and excluding liquids). Most of these compounds are not stable at our chosen T and P . We derive our fiducial
list of compounds starting from the list reported in Visscher & Fegley (2013) and the set of compounds in Pignatale
et al. (2011). Complex organics have been excluded from calculations as their chemistry is driven more by kinetics
rather than thermodynamic equilibrium. C-graphite is taken as representative of the main carbon condensates together
with Fe3C, Fe2C and SiC. Calcium and aluminium refractory species, all main oxides and main silicates (Mg and Fe
silicates) have been taken into account. Sulfides are included as well as water-vapour and water-ice. We report the
complete list of considered species in Table 2. The following nomenclature will be used: olivine (forsterite, Mg2SiO4,
and fayalite, Fe2SiO4), pyroxene (enstatite, MgSiO3, and ferrosilite, FeSiO3), plagioclase (anorthite, CaAl2Si2O8
and albite, NaAlSi3O8), melilite (gehlenite, Ca2Al2SiO7, and akermanite, Ca2MgSi2O7), fassaite (Ca-Tschermak,
CaAl2SiO6, and diopside, CaMgSi2O6), spinel (MgAl2O4 and FeAl2O4), magnesiowustite (MgO and FeO), sulfide
(FeS, MgS and CaS), metal (Fe, Al and Zn). Only the endmembers of each solids solution are considered and no
predictions of intermediate compositions are made.
To summarize, we calculate the thermodynamic equilibrium for each of the considered cases in table 1 at the given
temperature (T = 2000 K) and pressure (P = 10−4 bar). The resulting gas phase of Mars plus the gas phase of the
selected impactor will constitute the gas mixture from which the dust will condense. The material that is not in the
2 We take the elemental abundances from Asplund et al. (2009) for the given set of elements. Please note that He is not included in the
system and, as a consequence, the abundance of H raises to ∼99% of the total.
AASTEX Dust composition from impactors
5
Table 1. Elemental abundances (mol%) for single impactors. Abundances of the solar photosphere (Sun) are also shown.
Mars
CV
CI
EH
comet
Sun
Element
Abundances (mol%)
Al
C
Ca
Fe
H
K
Mg
Na
O
S
Si
Ti
Zn
1.250E+00
1.356E+00
4.668E-01
7.596E-01
7.000E-02
2.82E-04
0.000E+00
9.747E-01
3.902E+00
8.427E-01
1.137E+01
2.69E-02
9.300E-01
9.911E-01
3.362E-01
5.367E-01
6.000E-02
2.19E-04
5.440E+00
8.797E+00
4.773E+00
1.314E+01
5.200E-01
3.16E-03
0.000E+00
5.808E+00
2.906E+01
0.000E+00
5.464E+01
99.9
4.800E-02
1.658E-02
2.098E-02
5.177E-02
0.000E+00
1.07E-05
1.632E+01
1.247E+01
5.845E+00
1.104E+01
9.900E-01
3.98E-03
7.000E-01
3.001E-01
3.121E-01
7.484E-01
1.000E-01
1.74E-04
5.815E+01
5.157E+01
4.491E+01
4.724E+01
2.834E+01
4.89E-02
0.000E+00
1.434E+00
2.695E+00
4.577E+00
7.100E-01
1.32E-03
1.673E+01
1.624E+01
7.656E+00
2.104E+01
1.830E+00
3.23E-03
4.000E-02
4.280E-02
1.285E-02
2.379E-02
0.000E+00
8.90E-06
2.000E-04
3.709E-03
6.988E-03
9.674E-03
0.000E+00
3.63E-06
Mg/Si
Fe/O
Fe/Si
(Fe+Si)/O
C/O
H/O
0.98
0.09
0.33
0.38
0.00
0.00
0.77
0.17
0.54
0.49
0.02
0.11
0.76
0.11
0.62
0.28
0.09
0.65
0.52
0.28
0.62
0.72
0.02
0.00
0.54
0.02
0.28
0.08
0.40
1.93
1.25
0.06
0.98
0.13
0.54
2041
gas phase will form the melt from which the solids will form. To derive the dust composition we then proceed to
the computation of the condensation sequence in the interval of temperatures of 150 < T (K) < 2000 with a constant
pressure of P = 10−4 bar. To derive the solids composition we compute the CIPW-norm.
In order to test our thermodynamic model we also run equilibrium calculation using the solar abundances in Table 1
and compare the results with previous calculations available in the literature. Results of the test and a brief discussion
are presented in Appendix A.
3. RESULTS
Table 3 shows the elemental abundances (mol%) of the gas mixture that results from equilibrium calculation at
T = 2000 K and P = 10−4 bar of Mars plus the considered impactor (Mars+Mars, Mars+CV, Mars+CI, Mars+EH,
Mars+comet). These abundances are used as input to compute the condensation sequence. Table 4 show the oxides
budget of different melt mixtures in case of complete equilibration between Mars and given impactors. These budgets
are used to compute the CIPW-norm.
3.1. Dust from condensing gas
Figure 2 shows the dust distribution for all the considered impactors in mol% (being 100 gas+dust) as a function of
temperature. From left to right and from top to bottom, the different cases are ordered with decreasing Fe/O ratio of
the initial gas mixture (see Table 3).
Mars+CV impact results in large quantities of metallic iron, FeS and SiO2. Small amount of pyroxene (enstatite
(MgSiO3) and ferrosilite (FeSiO3)), ∼ 1 mol%, is distributed all along the temperature range. At T ∼ 700 K, we
see the appearance of Fe2C and C (graphite). Similarly to Mars+CV, the Mars+EH impact shows large quantities of
metallic iron, FeS and SiO2. Moreover, we do see small percentage of Si, MgS and SiC. Traces of pyroxenes are seen
at high temperatures only.
6
Pignatale et al.
Table 2. Complete list of gas and dust species in the equilibrium calculations.
Gas
Al(g) Al2O2(g) Al2O3(g) AlO(g) AlO2(g)
C(g) Ca(g) CaO(g) CH4(g) CO(g)
Fe Fe(g) FeO(g) FeS(g)
H(g) H2(g) H2O(g) H2S(g) HS(g)
K(g) K2(g) K2O(g) KO(g) Mg(g) MgO(g)
Na(g) Na2(g) Na2O(g) NaO(g)
O(g) O2(g) OH(g)
S(g) Si(g) SiO(g) SiO2(g)
Ti(g) TiO(g) TiO2(g)
Zn(g) ZnO(g)
Dust
Al2O3
C
Ca2Al2SiO7 Ca2MgSi2O7 Ca2SiO4 CaAl12O19 CaAl2O4 CaAl2Si2O8
CaAl2SiO6 CaAl4O7 CaMgSi2O6 CaO CaS CaSiO3 CaTiSiO5
Fe Fe2C Fe2O3 Fe2SiO4 Fe3C Fe3O4
FeAl2O4 FeO FeSiO3 FeTiO3
H2O
K K2O K2Si4O9 KAlSi2O6 KAlSi3O8 KAlSiO4
Mg2Al4Si5O18 Mg2SiO4 Mg2TiO4 MgAl2O4 MgO MgS MgSiO3 MgTi2O5 MgTiO3
Na2O Na2SiO3 NaAlSi3O8
Si SiC SiO2
TiO2
Zn Zn2SiO4 Zn2TiO4 ZnO ZnSiO3
The Mars+Mars impact produces several oxides such as FeO, Fe3O4, metallic iron and volatiles such as Na2O,
Na2SiO3. Traces of olivine (forsterite (Mg2SiO4) and fayalite (Fe2SiO4)) are present at high temperature.
Mars+CI impact returns iron-rich olivine such as fayalite (Fe2SiO4), then FeO, Fe3O4, Fe2O3, FeS and smaller
amount of SiO2. At lower temperature we see the condensation of C and H2O. The dust from Mars+comet impact is
mainly made of pyroxene, SiO2 and FeS. Mars+comet impact is that one that produces, as expected, a large amount
of water ice together with solid carbon.
Figure 3 shows the condensation sequence for the more volatiles species. All the considered cases return a very similar
behaviour as these volatiles are less effected by the changes of other elemental ratios. Na(g) has higher condensation
temperature than K(g) and Z(g) is the last one to condense. Na2SiO3, K2Si4O9 and Zn2SiO4 are the main respective
condensates, together with Zn in case of Mars+CV, Mars+EH, and Zn and K for Mars+Mars.
3.2. Solids from cooling melts
Table 5 reports the resulting CIPW-norm if complete equilibration between the melt belonging to Mars and to the
impactor occurs (see Table 4). To establish the reliability of our CIPW-algorithm, we performed calculation using
the BSM and compare our result with that in Ronnet et al. (2016) finding a very good agreement (see the last two
columns in Table 5).
AASTEX Dust composition from impactors
7
Table 3. Elemental abundances (mol%) of the gas mixture which is released after the impact assuming T = 2000 K, P =
10−4 bar, different types of impactors, and equal contribution between Mars and the considered impactor.
Gas mixture
+Mars
+CV
+CI
+EH
+comet
Element
Abundances (mol%)
Al
C
Ca
Fe
H
K
Mg
Na
O
S
Si
Ti
Zn
8.059E-06
1.382E-05
5.058E-06
2.336E-05
1.412E-05
0.000E+00
3.762E+00
5.791E+00
2.060E+00
1.141E+01
2.348E-05
4.508E-05
9.962E-06
8.897E-05
1.804E-04
1.974E+01
2.972E+01
6.652E+00
3.117E+01
8.041E-01
0.000E+00
2.241E+01
4.314E+01
0.000E+00
5.481E+01
3.360E+00
2.489E-01
1.008E-01
2.423E-01
4.808E-02
3.539E-01
4.512E-01
1.000E-01
8.458E-01
9.982E-01
4.905E+01
3.859E+00
1.502E+00
3.543E+00
8.026E-01
2.608E+01
2.501E+01
3.695E+01
2.775E+01
2.858E+01
0.000E+00
5.519E+00
4.000E+00
1.120E+01
7.121E-01
1.146E+00
8.663E+00
1.735E+00
2.310E+01
1.825E+00
2.663E-01
3.370E-01
2.347E-02
6.733E-02
3.844E-03
1.051E-02
1.505E-02
1.158E-02
2.398E-02
2.007E-04
Mg/Si
Fe/O
Fe/Si
(Fe+Si)/O
C/O
H/O
0.31
0.76
17.23
0.80
0.00
0.00
0.05
1.19
3.43
1.53
0.15
0.90
0.06
0.18
3.83
0.23
0.16
1.17
0.04
1.12
1.35
1.96
0.07
0.00
0.55
0.03
0.44
0.09
0.40
1.92
Table 4. Oxide composition (wt%) of the melt that results after the impact assuming T = 2000 K, P = 10−4 bar and different
types of impactor. Total equilibration between Mars and impactor is also assumed. The Mg/Si and Fe/O ratios are the elemental
mole ratios.
+Mars +CV +CI +EH +comet
Al2O3
CO2
CaO
FeO
H2O
K2O
MgO
Na2O
SiO2
TiO2
ZnO
SO3
Mg/Si
Fe/O
2.96
0.00
2.40
3.55
0.00
2.87
3.06
0.00
2.47
3.02
0.00
2.41
3.10
0.00
0.28
17.16
12.55
14.42
12.35
17.06
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
30.60
30.88
31.15
32.04
31.09
0.00
0.00
0.00
0.00
0.00
46.75
49.35
48.80
50.10
45.11
0.13
0.00
0.00
0.98
0.09
0.80
0.00
0.00
0.93
0.06
0.10
0.00
0.00
0.95
0.07
0.09
0.00
0.00
0.95
0.06
3.37
0.00
0.00
1.03
0.09
8
Pignatale et al.
Figure 2. Condensation sequences for major dust species (mol%) that result from the gas mixtures in Table 3. Note the
changes of colours when different compounds are considered.
AASTEX Dust composition from impactors
9
Figure 3. Condensation sequences for K, Zn, and Na compounds (mol%) that result from the gas mixtures in Table 3. Sodium
compounds for Mars+Mars were included in Fig. 2 as, in that case, they represent major species. Note the changes of scales in
the y-axis.
10
Pignatale et al.
Table 5. Resulting CIPW-norm of the melt phase. Calculations for the BSM are also performed to compare the resulting
CIWP-norm with values derived by Ronnet et al. (2016).
+Mars +CV +EH +CI +COMET
BSM BSM (Ronnet et al. 2016)
Anorthite
Diopside
8.08
3.08
9.69
3.63
8.24
2.97
8.35
3.13
1.39
0.00
3.16
6.89
Pyroxene
43.41
55.48
57.58
52.35
54.97
21.03
Albite
Orthoclase
Olivine
Ilmenite
Corundum
Anorth+Alb
Oli/Pyr
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
0.00
8.29
0.65
45.19
29.68
31.05
35.98
34.66
58.50
0.25
0.00
0.00
1.04
1.52
0.00
0.00
0.53
0.17
0.00
0.00
0.62
0.19
0.00
0.00
0.68
6.40
2.59
0.00
0.63
0.27
0.00
(11.45)
2.78
6.97
21.29
0.66
59.22
0.00
0.00
11.59
2.78
The resulting solids will be generally characterized by pyroxene3 and olivine, with the former in larger abundances,
except for the Mars+Mars case for which olivine is slightly more abundant. Although our selected impactors have
initially different chemical composition, the resulting CIPW-norm is quite similar for all cases. It is interesting to
note that diposide (CaMgSi2O6) is not predicted for a cometary impactor, while corundum (Al2O3) is a tracer of that
impact. Enstatite and forsterite will be largely stable and common compounds for all the considered cases. Albite
(NaAlSi3O8) and orthoclase (KAlSi3O8) are not present in the solids because Na and K are totally vaporized after
the impact (see Tables 3 and 4).
4. DISCUSSION
4.1. Dust composition
Our calculations clearly show different behaviour when compared with the classical condensation sequence with a
solar composition (see Fig. 5). One of the main reasons is the amount of H, C and O in our systems that is very
different from the solar values. Moreover, in our calculations the amount of Fe, Mg, and Si is of the same order of
magnitude as O. This is not the case for the Solar Nebula where H is predominant, C is comparable with O and Fe,
Mg, Si are orders of magnitude smaller than O (Asplund et al. 2009). Here we try to qualitatively understand our
results and emphasize the differences from the well known condensation sequence of the Solar Nebula.
The stability of forsterite (Mg2SiO4) and enstatite (MgSiO3) is driven by the Mg/Si ratio where higher Mg/Si ratios
(> 1) favour forsterite while lower Mg/Si ratios (< 1) favour enstatite (Ferrarotti & Gail 2001). However, at very
high temperature, forsterite can still be the first magnesium silicates that condenses out before being converted in
enstatite (Ferrarotti & Gail 2001). From Table 3 we see that the Mg/Si is well below 1 in all cases and, as expected
the dust is generally enstatite-rich (see Fig. 2). The excess of Si that is not consumed in the magnesio-silicates will
then be bound with O to form stable SiO2. Generally SiO2 tends to be more stable than iron-oxides (see for example
ellingham diagrams in DeHoff (1993)). Fe and SiO2 can then react to form iron-rich silicates. If oxygen is still available
for reaction it will start to bind iron to form iron-oxides. If there is lack of oxygen, iron will be mainly in the metallic
form. The presence of sulfur further modifies the expected composition as sulfidation of Fe occurs.
As a consequence, looking at the elemental ratios reported in Table 3 the behaviours found in Fig. 2 become clearer.
Let us consider the two extreme cases of the Fe/O ratio, Mars+comet (Fe/O=0.03) and Mars+CV (Fe/O=1.19). In
the case of Mars+comet we have Mg/Si=0.55. As such we expect the oxygen to form mainly enstatite MgSiO3. Then
we expect the appearance of SiO2 as there is Si in excess with Si more abundant than Fe (Fe/Si=0.44). As Fe/O=0.03
and (Fe+Si)/O=0.09, there is a large reservoir of oxygen to oxidise the iron which, indeed, is found in ferrosilite
FeSiO3.
3 hypersthene in Ronnet et al. (2016).
AASTEX Dust composition from impactors
11
On the other hand, let us focus to the Mars+CV case. Here the small amount of Mg will form magnesium-silicates,
then the excess of Si will form SiO2 and FeSiO3. The Fe/Si=3.43 tells us that there is iron in excess compared to Si.
The Fe/O=1.19 and (Fe+Si)/O=1.53 also tell us that there is not enough oxygen available to oxidise all the iron. As
a consequence, we expect to see a certain amount of iron to be stable in its metallic form, and FeS given the presence
of sulfure in the mixture. Sulfur is also present in the case for Mars+EH, Mars+CI and Mars+Comet. If large amount
of sulfur is available, as in the the Mars+EH impact, MgS also becomes a stable sulfide.
Interesting cases are also the Mars+Mars and Mars+CI. Here, the high Fe/Si ratios but low Fe/O ratios return large
amount of several iron oxides: there is no enough Si to form large amount of iron-silicates, thus the O binds directly
the iron in several iron-oxides. The Mars+EH impact clearly shows the effect of the sulfide-rich impactor given the
presence of MgS together with FeS.
Our calculations show that different impactors result in dust with traceable different composition. This open the
possibility to identify the individual composition of the impactor from the determination of the dust composition of
Phobos and from the samples collected by the MMX mission.
In conclusion, dust from different bodies will be characterized by different i) degrees of iron-oxidation, ii) presence of
iron-silicates and/or iron-oxides, iii) amount of sulfides, iv) amount of carbon and ice. All the condensation sequences
return a generally poor content of olivine and pyroxene, with a preference of the latter. A qualitative analysis of
different elemental ratios can then be useful to derive the chemistry of impactors that are not considered in this work.
4.1.1. Carbon, water ice and other volatiles
In our calculations we see the appearance of solid C in the case of Mars+CV, Mars+CI and Mars+comet, while SiC
is the most stable C-bearing compound in the case of Mars+EH . Mars+CV, Mars+CI and Mars+comet have carbon
and hydrogen in the gas mixture, while in the Mars+EH case there is carbon only.
The carbon-to-oxygen ratio (C/O) is an important parameter that determines the presence of solid carbon, water
vapour and other oxides. At high temperature CO(g) will consume all the C available before allowing the formation of
water vapour (Larimer 1975). This has strong implication in the formation of complex organics and water (if hydrogen
is present in the system).
The chemistry of carbon cannot be totally determined by thermodynamic equilibrium. The behaviour of CO(g)
in a H2O(g)-H2(g) gas in the temperature range of 100 < T (K) < 700 K is ruled by kinetics and the environmental
condition. The classical transition around T ∼ 700 K where CO(g) is transformed in CH4 can be described, for
(cid:10) CnH2n+2 + nH2O. This can be noticed in Fig.5(right) where
example, by a reaction of the type nCO + (1 + 2n)H2
n = 1 and the reaction is CO(g) + 3H2(g) (cid:10) CH4(g) + H2O(g).
In fact, this reaction is just a first simplified transcription of the many Fischer-Tropsh-like reactions that can occur in
this temperature range and at different H2/CO ratios (Sekine et al. 2006; Nuth et al. 2008). Fischer-Tropsh processes
produce complex organics on the surface of dusty grains in the presence of the right catalyst. There are different
catalysts with their own properties, but usually, in astrophysics context, the iron-based Fischer-Tropsh is the most
considered because the abundances of this element in the Solar Nebula (Fegley 1988; Asplund et al. 2009).
There are numerous competitive reactions that determine the rate and the production of organics (Fegley 1988) and
several theoretical models have been largely described in material science (see for example the work of Zimmerman
& Bukur (1990)). However, the resulting amount of organics is extremely difficult to calculate theoretically when
an astrophysical environment is considered. This comes from the large uncertainty in determining, for example, the
amount and the surface of catalyst available.
Nevertheless, the possible pathways to the formation of carbon-rich material are vastly more numerous.
In our
Mars+comet case, for example, the reaction CO(g) + H2(g) (cid:10) C + H2O(g) becomes active as well. This is clearly
seen in figure 4 where H2(g) and COg are depleting as CH4(g), H2O(g) and C become more stable. In this work we
do not perform any kinetics calculation and, since we have only few carbon-solids in our list, we can only suggest
that the presence of C and H in Mars+CI, Mars+Comet and Mars+CV impacts can produce complex organics and
carbon-enriched dust. The case of Mars+CV is then extremely interesting because there may be enough metallic iron,
together with carbon, to enhance the production of complex organics.
Equilibrium calculations return an efficient evaporation of the carbon rich dust present in the impactor. However,
the rate of vaporisation will be driven by the physical and chemical properties of the carbon species. For example,
carbon rich insoluble organic material (IOM), if present, could survive the impact as it is refractory (Pizzarello et al.
12
Pignatale et al.
Figure 4. Thermodynamic calculations in the +comet case shows a possible pathway for the condensation of graphite. In fact,
together with the well known transformation of CO(g) in CH4(g) (CO(g) + 3H2(g) (cid:10) CH4(g) + H2O(g)), we do see the following
reaction: CO(g) + H2(g) (cid:10) C + H2O(g) that occurs at temperature lower than T=600 K.
2006). This could reduce the amount of carbon released in the gas phase. Nevertheless, the presence of carbon in the
MMX samples in the form of new condensates and/or in IOM will still be tracer of the nature of the impactor.
In section 3 we pointed out that the gas-mixture is volatile-enriched. Figure 3 shows that K-, Na-, and Zn-silicates
are stable as the temperature drops. Mono atomic Zn is also predicted for the Mars+EH and Mars+Mars case. There
are no dramatic differences between the considered impactors when K, Na and Zn are taken into account.
In conclusion, the condensing dust will be volatile (and in some cases, also carbon) enriched compared to the
solids that result from the cooling melts. The presence of volatile-rich dust in MMX samples will thus indicate that
vaporization followed by condensation had occurred and no volatiles left the system. Moreover, since Na, K and
Zn condense at different temperature, their presence/absence could return information on the temperature at which
aggregation of a given sample occurred.
Mars+CI and Mars+comet are the only cases for which, at low temperature, we see condensation of water vapour
into ice. The presence of ice will favour secondary alteration of the dust allowing, for example, the formation of
phyllosilicates (Bischoff 1998).
4.2. Solids composition
In table 5 we reported the resulting CIPW-norm of the solids if complete equilibration between Mars and impactor
occurred. The composition of the solids generally comprises olivine (forsterite and fayalite) and pyroxene (enstatite
and ferrosilite). Ronnet et al. (2016) calculated the CIPW normative mineralogy for a Mars, Moon and IDP like
impactor. They found that the resulting composition of a mars-like impactor would be olivine and pyroxene rich. In
particular their CIPW-norm for Mars is characterized by high olivine content (olivine/pyroxene > 1).
olivine/pyroxene ∼ 1, whereas the other cases return a olivine/pyroxene < 1.
Our results for the Mars+Mars case show that the solids (deprived of all the vaporized material) will have an
There are no dramatic differences between the solids that results from different impactors (except for the aforemen-
tioned corundum in the comet case). It is interesting to note that solids that result form cooling melts do not show
as much as variation in their composition compared to the dust. The resulting composition of solids appears, thus, to
contain less information on the origin of the impactor compared to the large quantities of clues that can be derived from
condensed dust. Nonetheless, the composition of melts can be affected by different cooling conditions, microgravity
and gas fugacities. Nagashima et al. (2006) and Nagashima et al. (2008) performed laboratory experiments on cooling
forsterite and enstatite melts. They found that different cooling rates and microgravity can alter and even suppress
crystallisation only allowing the formation of glass material. Further experimental investigations are already planned
in order to derive predictions of the composition of the cooling melts.
AASTEX Dust composition from impactors
13
On the physical point of view, condensates and solids from melts may be distinguished by their different crystalline
structure, microporosity, zoning, interconnections between different phases. Indeed, the resulting physical properties
of dust from gas and solids from melts are determined by many factors (Nishinaga 2014). Furthermore Hyodo et al.
(2017a) showed that while the size of dust would be in the order of 0.1∼10µm, solids from melts can reach 1∼10m in
size and then they can be grinded down to ∼100µm. We could, thus, expect to find different size distributions when
dust and solids are compared.
4.3. Infrared spectra of Phobos
Giuranna et al. (2011) presented a detailed investigation on the possible composition of the dust and rocks present
on the Phobos surface. They suggested that the "blue" part of Phobos is consistent with phyllosilicates while the
"red" region is compatible with the presence of feldspar. No bulk chondrite compositions are able to reproduce the
current observation (Giuranna et al. 2011).
Phyllosilicates are not product of condensations but derive from secondary alterations of silicates (Bischoff 1998)
and, as a consequence, they are not predictable with our calculations, although we do have all the dust (silicates)
at the base of their formation. The major feldspar compounds are orthoclase (KAlSi3O8), albite (NaAlSi3O8), and
anorthite (CaAl2Si2O8). Not all of them are compatible with our model as Na and K are separated from Al after the
impact and others are the predicted stable compounds. On the other hand we do see the formation of anorthite (see
Table 5).
Nevertheless Giuranna et al. (2011) stressed that more fine modelling is needed as a mixture of different materials
made of fine grains could also produce the observed trends. This becomes important as previous modelling focused on
the analysis of the spectral properties of "external" objects (such as different types of asteroids) to match the observed
spectra, as the capture scenario suggests. The impact-generated scenario imposes to re-think this approach.
Ronnet et al. (2016) analysed the resulting composition of the melts generated by different impactors in order to find
a match with the observations. They concluded that more than the melt, the gas-to-dust condensation in the outer
part of an impact-generated disk could be able to explain the Phobos and Deimos spectral properties. This further
suggests that our derived dust may play an important part in producing the observed trends. Moreover, since the
Hyodo et al. (2017a) disc model shows that melt would be mixed with the gas, the combined effect of dust and solids
should be taken into account.
In this section we try to predict the effects of our mixed material (dust plus solids) on the infrared spectra. In the
introduction we presented possible mechanisms listed by Giuranna et al. (2011) that could be able to reproduce the
observed trend in the VIS-NIR. Here we recall them and compare with our results:
i) Low percentage of iron-rich olivine and pyroxene can reduce the spectra. Our resulting dust mixtures generally
have a very low concentration (∼ 1/ mol% ) of iron-rich olivine (fayalite, Fe2SiO4) and pyroxene (ferrosilite, FeSiO3).
Only the high temperature region of Mars+CI shows a larger amount of fayalite (see Fig. 2).
ii) A mixture of opaque material (metal iron, iron-oxide, and carbon) reduce the emissivity. We do have a metallic
iron-rich dust that results from several impactors (Mars+CV, Mars+EH, and Mars+Mars only at low temperature).
Carbon dust is also seen in our calculations (Mars+CV, Mars+CI, Mars+comet). Moreover, together with Fe, iron
sulfide (FeS), that we see in Mars+CV, Mars+EH, Mars+CI, Mars+comet, is opaque and featureless in the NIR , but
may be recognizable in the MID-IR (Wooden 2008; Henning & Meeus 2011).
assemblage, can have the characteristics suggested by Giuranna et al. (2011).
iii) Quenched material lacks of perfect crystalline structure and, thus, reflectance. Solids from the melts, in our final
iv) The reflectance of fine grains is reduced. The average size of the condensed dust is in the order of 0.1∼10µm
Hyodo et al. (2017a).
Our proposed model of Phobos as a result of accretion of dust from gas condensation and solids from melts, together
with our derived chemical composition looks promising when discussing spectra. Nevertheless there are some aspects
that needs further investigation: i) it is important, at this point, to derive the MIDIR spectra of our propose mixtures,
and then ii) estimate the effect of space weathering on it and on the resulting albedo. These points can be set as main
topic for future works.
In this work we assume thermodynamic equilibrium (where all the reactions rates are much shorter than the disk
cooling timescale) and mass conservation. All the material is available for reaction until the equilibrium is reached at
4.4. Limitations
14
Pignatale et al.
any given temperature. This may be not always the case. As dust condenses out from the gas, it can be subject to
different drag forces and it can separate from the current environment. This process can lead to the so called "frac-
tionated" condensation sequence. If this is the case, a given dust grain can become representative of the temperature,
pressure, gas mixture and dust species that were present when its condensation occurred.
Moreover, dust from secondary condensations (from a now fractionated gas) may then form. These fractionated and
incomplete condensation sequences have been the subject of several studies (Hutson & Ruzicka 2000; Blander et al.
2009; Pignatale et al. 2016) which all show that different pathways of condensation can depart from the main line. For
example, Pignatale et al. (2016) showed that, starting with a gas of solar composition, a mixture of enstatite-rich and
SiO2-rich dust can be produced in case of systematic sharp separation between dust and gas. SiO2 is a condensate
that is not predicted ("incompatible") when solar abundances are considered.
In this work we do not perform fractionated condensation sequences as numerous are the possible pathways. However,
in the same way, the presence of some "incompatible" condensates together with "predictable" dust in the same MMX
sample may point out to incomplete or secondary condensation.
As mentioned, Hyodo et al. (2017a) demonstrated that there will be likely no complete equilibration between the
melts of Mars and the melts of the impactor. What is likely to occur is a wide spectrum with different degrees of
equilibration. A random sampling of solids may, thus, show material that come from Mars, the impactor or several
degrees of mixing.
In our calculations we kept the pressure constant and fixed at P = 10−4 bar as in Hyodo et al. (2017a). In general,
lower pressures (in order of magnitude) move the condensation of the dust toward lower temperatures (Yoneda &
Grossman 1995; Gail 1998). In this case, for our given temperature, more material could vaporize and go into the
gas-phase. Increasing the pressure (in order of magnitude) has the opposite effect as the condensation temperature
increases. As a consequence, we could observe different amount of Fe, Mg, and Si moving to the gas phase. Changes
in disk pressure may also occur if large amount of volatiles are injected in the system after the impact. This could
be the case of the Mars+comet impact where the release of H2O(g) and CO(g) could change the total pressure in
the disk increasing it, or when a water-rich Mars is considered (Hyodo et al. 2017a). Observed deviations from the
predicted trends could then be associated to strong variation in the pressure in the Mars' Moons formation region of
the disk or, in fact, to a radial gradient of temperature and pressure in the disk. As reference for future experimental
work we report in appendix B (see Fig.6) the partial pressures of the major gas component for Mars+CI, Mars+CV,
and Mars+comet impacts. These are the impacts that produce the larger amount of gas such as H2O(g) and CO(g).
These values can be used to set up the conditions in which experiments can be performed.
5. CONCLUSIONS
In this work we used thermodynamic equilibrium calculation to investigate the chemical composition of dust (from
condensing gas) and solids (from cooling melt) as the building blocks of Phobos and Deimos in the impact-generated
scenario with the thermodynamic conditions of Hyodo et al. (2017a).
We found that dust and solids have different chemical and physical properties. Dust carries more information on the
impactor than the solids. Our results show that it would be possible to distinguish from different types of impactors
as each case returns several unique tracers in the dust: a Mars+CV has large quantities of metallic iron, SiO2, iron
sulfides and carbon; Mars+Comet has pyroxenes and the largest carbon and ice reservoir. Mars+EH impact has dust
with high metallic iron content, SiO2, sulfides (FeS and MgS) and traces of SiC. Impact with Mars-like objects returns
several iron-oxides, and the dust in Mars+CI has iron-oxides, water ice and carbon.
The presence/absence of metallic iron, iron-silicates, iron-oxides, sulfides, carbon and water ice can be considered as
clues of different impactors. Deviations from the derived compositions can be then ascribed to fractionated conden-
sation sequences and/or strong variations in the disk pressure and/or impactors with different elemental composition
than investigated in this study.
The giant impact scenario imposes to re-think the dust modelling for the infrared spectra, as Phobos, in this case,
would be made of a complex mixture of dust and solids and not of a pre-built object as the capture scenario suggests.
A qualitative analysis suggests that our derived composition of dust and solid can be compatible with the characteristic
of the Phobos VIS-NIR spectra.
In conclusion, the proposed scenario of Phobos as the result of accretion of dust and solid in an impact-generated
disk can reconcile with both the dynamical and spectral properties of the Mars' moon. Our dust tracers can be then
used in the analysis of the samples returned by the JAXA's MMX mission.
AASTEX Dust composition from impactors
15
Figure 5. Condensation sequence for major solids (left) and gas (right) starting with an initial gas with solar composition.
There is a very good agreement between these results and those reported in Pignatale et al. (2011) and reference therein.
Software: HSC (v8; Roine (2002))
The authors wish to thank the anonymous referee for their comments and suggestions that let us investigate our
assumptions in more details which improved the manuscript. The authors wish to acknowledge the financial support
of ANR-15-CE31-0004-1 (ANR CRADLE), INFINITI (INterFaces Interdisciplinaires Num`erIque et Th`eorIque), Uni-
vEarthS Labex program at Sorbonne Paris Cite (ANR-10- LABX-0023 and ANR-11-IDEX-0005-02). PR has been
financially supported, for his preliminary contribution to this work, by the Belgian PRODEX programme managed
by the European Space Agency in collaboration with the Belgian Federal Science Policy Office. SC, RH, and HG
acknowledge the financial support of the JSPS-MAEDI bilateral joint research project (SAKURA program). HG also
acknowledges JSPS KAKENHI Grant Nos. JP17H02990 and JP17H06457, and thanks the Astrobiolgy Center of the
National Institutes of Natural Sciences (NINS).
APPENDIX
A. SOLAR CONDENSATION SEQUENCE
Condensation sequences calculated with solar abundances are very distinctive and characterized by smooth transi-
tions between refractories, main silicates and metallic iron and low temperature material (such as sulfides) (Larimer
1979; Yoneda & Grossman 1995; Lodders 2003; Pignatale et al. 2011).
Fig. 5 reports the condensation sequence calculated with our thermodynamic system and the solar abundances listed
in Table 1 in the temperature range of 150 < T (K) < 2000 and pressure P = 10−4 bar. The small fraction of solids that
are stable at high temperature (T > 1400 K) are the refractories calcium-alluminum silicates while at T ∼ 1400 K iron,
forsterite and enstatite condense. At lower temperature we see the formation of troilite (FeS) and fayalite (Fe2SiO4).
CO(g), H2O(g) and SiO(g) are the main O-binding gaseous species. At T ∼ 700 K we see the conversion of CO(g) to
CH4, while H2S is the main sulfur-binding gas until the condensation of troilite.
These results are in very good agreement with previous works (Larimer 1979; Yoneda & Grossman 1995; Lodders
2003) on the solar condensation sequence and make us confident on our built system.
B. PARTIAL PRESSURES OF MAIN GASEOUS COMPOUNDS
In figure 6 we report the partial pressures for the main gaseous species as they result from our equilibrium calculations.
The chosen case are Mars+CI, Mars+CV and Mars+comet. These are the impacts that introduce H, C, and O in the
system. In our calculations ideal gas are considered. As a consequence their partial pressure can be taken as a proxy
for their fugacity. These values can be used to set the initial gas environment in further experimental studies.
16
Pignatale et al.
Figure 6. Partial pressure for major gas in the Mars+CI, Mars+CV and Mars+comet impacts. The total pressure of the
system is P = 10−4 bar.
REFERENCES
Asplund, M., Grevesse, N., Sauval, A. J., & Scott, P. 2009,
-. 1990c, Icarus, 84, 315
A&ARA, 47, 481
Bischoff, A. 1998, Meteoritics and Planetary Science, 33,
1113
Blander, M., Pelton, A. D., & Jung, I.-H. 2009, Meteoritics
and Planetary Science, 44, 531
Bond, J. C., O'Brien, D. P., & Lauretta, D. S. 2010, ApJ,
715, 1050
Bottke, W. F., & Andrews-Hanna, J. C. 2017, Nature
Geoscience, 10, 344348
Cloutis, E. A., Smith, D. G. W., Lambert, R. S. J., &
Gaffey, M. J. 1990d, Icarus, 86, 383
Craddock, R. A. 2011, Icarus, 211, 1150
Cross, W., Iddings, J. P., Pirsson, L. V., & Washington,
H. S. 1902, The Journal of Geology, 10, 555
DeHoff, R. 1993, Thermodynamics in Materials Science
(McGraw-Hill International Editions)
Ebel, D. S. 2006, Condensation of Rocky Material in
Brasser, R., & Mojzsis, S. J. 2017, Geophys. Res. Lett., 44,
Astrophysical Environments, ed. D. S. Lauretta & H. Y.
5978
McSween, 253–277
Brasser, R., Mojzsis, S. J., Matsumura, S., & Ida, S. 2017,
Ebel, D. S., & Grossman, L. 2001, GeoCoA, 65, 469
Earth and Planetary Science Letters, 468, 85
Burns, R. G. 1993, Mineralogical applications of crystal
field theory, Vol. 5 (Cambridge University Press)
Citron, R. I., Genda, H., & Ida, S. 2015, Icarus, 252, 334
Cloutis, E. A., Gaffey, M. J., Smith, D. G. W., & Lambert,
R. S. J. 1990a, J. Geophys. Res., 95,
doi:10.1029/JB095iB06p08323
-. 1990b, J. Geophys. Res., 95, 281
Fegley, Jr., B. 1988, in Origins of Solar Systems, ed. J. A.
Nuth & P. Sylvester
Ferrarotti, A. S., & Gail, H.-P. 2001, A&A, 371, 133
Gail, H.-P. 1998, A&A, 332, 1099
Gail, H.-P., & Sedlmayr, E. 1999, A&A, 347, 594
Giuranna, M., Roush, T. L., Duxbury, T., et al. 2011,
Planet. Space Sci., 59, 1308
AASTEX Dust composition from impactors
17
Henning, T., & Meeus, G. 2011, Dust Processing and
Mineralogy in Protoplanetary Accretion Disks, ed.
P. J. V. Garcia, 114–148
Hesselbrock, A. J., & Minton, D. A. 2017, Nature
Geoscience, 10, 266
Huebner, W. F., & Boice, D. C. 1997, Polymers and Other
Macromolecules in Comets, ed. P. J. Thomas, C. F.
Chyba, & C. P. McKay (New York, NY: Springer New
York), 111–129
Hutson, M., & Ruzicka, A. 2000, Meteoritics and Planetary
Science, 35, 601
Hyodo, R., Genda, H., Charnoz, S., & Rosenblatt, P. 2017a,
ApJ, 845, 125
Pasek, M. A., Milsom, J. A., Ciesla, F. J., et al. 2005,
Icarus, 175, 1
Pignatale, F. C., Liffman, K., Maddison, S. T., & Brooks,
G. 2016, MNRAS, 457, 1359
Pignatale, F. C., Maddison, S. T., Taquet, V., Brooks, G.,
& Liffman, K. 2011, MNRAS, 414, 2386
Pizzarello, S., Cooper, G. W., & Flynn, G. J. 2006, The
Nature and Distribution of the Organic Material in
Carbonaceous Chondrites and Interplanetary Dust
Particles, ed. D. S. Lauretta & H. Y. McSween, 625–651
Raymond, S. N., & Izidoro, A. 2017, Icarus, 297, 134
Rivkin, A. S., Brown, R. H., Trilling, D. E., Bell, J. F., &
Hyodo, R., Rosenblatt, P., Genda, H., & Charnoz, S. 2017b,
ApJ, arXiv:1711.02334
Plassmann, J. H. 2002, Icarus, 156, 64
Roine, A. 2002, User's guide, version, 5
Klima, R. L., Pieters, C. M., & Dyar, M. D. 2007,
Ronnet, T., Vernazza, P., Mousis, O., et al. 2016, ApJ, 828,
Meteoritics and Planetary Science, 42, 235
109
Larimer, J. W. 1975, GeoCoA, 39, 389
-. 1979, Icarus, 40, 446
Lodders, K. 2003, ApJ, 591, 1220
Lodders, K., & Fegley, B. 1997, in American Institute of
Physics Conference Series, Vol. 402, American Institute
of Physics Conference Series, ed. T. J. Bernatowicz &
E. Zinner, 391–423
Madhusudhan, N., Lee, K. K. M., & Mousis, O. 2012,
ApJL, 759, L40
Murchie, S. 1999, J. Geophys. Res., 104, 9069
Murchie, S. L., Thomas, P. C., Rivkin, A. S., & Chabot,
N. L. 2015, Phobos and Deimos, ed. P. Michel, F. E.
DeMeo, & W. F. Bottke, 451–467
Murchie, S. L., Britt, D. T., Head, J. W., et al. 1991,
Rosenblatt, P. 2011, A&A Rv, 19, 44
Rosenblatt, P., Charnoz, S., Dunseath, K. M., et al. 2016,
Nature Geoscience, 9, 581
Salisbury, J. W., & Walter, L. S. 1989, J. Geophys. Res.,
94, doi:10.1029/JB094iB07p09192
Sekine, Y., Sugita, S., Shido, T., et al. 2006, Meteoritics
and Planetary Science, 41, 715
Singer, R. B. 1981, J. Geophys. Res., 86, 7967
Visscher, C., & Fegley, Jr., B. 2013, ApJl, 767, L12
Wasson, J. T., & Kallemeyn, G. W. 1988, Philosophical
Transactions of the Royal Society of London Series A,
325, 535
White, W., Johnson, S., & Dantzig, G. 1958, J. Chemical
J. Geophys. Res., 96, 5925
Phys., 28, 751
Nagashima, K., Moriuchi, Y., Tsukamoto, K., Tanaka,
Wooden, D. H. 2008, Cometary Refractory Grains:
K. K., & Kobatake, H. 2008, Journal of Mineralogical
and Petrological Sciences, 103, 204
Interstellar and Nebular Sources, ed. H. Balsiger,
K. Altwegg, W. Huebner, T. Owen, & R. Schulz (New
Nagashima, K., Tsukamoto, K., Satoh, H., Kobatake, H., &
York, NY: Springer New York), 75–108
Dold, P. 2006, Journal of Crystal Growth, 293, 193
Nishinaga, T. 2014, Handbook of Crystal Growth:
Fundamentals (Elsevier)
Nuth, III, J. A., Johnson, N. M., & Manning, S. 2008,
ApJL, 673, L225
Yoneda, S., & Grossman, L. 1995, Geochimica
Cosmochimica Acta, 59, 3413
Zimmerman, W., & Bukur, D. 1990, Can. J. Chem. Eng.,
68, doi:10.1002/cjce.5450680215
|
1611.00051 | 2 | 1611 | 2017-03-21T20:40:30 | CO in distantly active comets | [
"astro-ph.EP"
] | Activity of most comets near the Sun is dominated by sublimation of frozen water, the most abundant ice in comets. Some comets, however, are active well beyond the water-ice sublimation limit of ~3 AU. Three bodies dominate the observational record and modeling efforts for distantly active comets: the long-period comet C/1995 O1 Hale-Bopp and the short-period comets (with Centaur orbits) 29P/Schwassmann Wachmann 1 and 2060 Chiron. We summarize what is known about these three objects emphasizing their gaseous comae. We calculate their CN/CO and CO2/CO production rate ratios from the literature and discuss implications. Using our own data we derive CO production rates for all three objects, in order to examine a correlation between gas production and different orbital histories and/or size. We find that orbital history does not appear to play a significant role in explaining 29P's CO production rates. 29P outproduces Hale-Bopp at the same heliocentric distance, even though it has been subjected to much more solar heating. Previous modeling work on such objects predicts that 29P should have been de-volatilized over a fresher comet like Hale-Bopp. This may point to 29P having a different orbital history than current models predict, with its current orbit acquired more recently. On the other hand, Chiron's CO measurements are consistent with it being significantly depleted over its original state, perhaps due to increased radiogenic heating made possible by its much larger size or its higher processing due to orbital history. Observed spectral line profiles are consistent with development and sublimation of icy grains at about 5-6 AU for 29P and Hale-Bopp, and this is probably a common feature in distantly active comets, and an important source of volatiles for all comets within 5 AU. In contrast, the narrow CO line profiles indicate a nuclear, and not extended, origin for CO beyond ~4 AU. | astro-ph.EP | astro-ph | CO in distantly active comets
M. Womack1, G. Sarid2, K. Wierzchos1
1University of South Florida, 2Florida Space Institute, University of Central Florida
Received
;
accepted
7
1
0
2
r
a
M
1
2
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
1
5
0
0
0
.
1
1
6
1
:
v
i
X
r
a
– 2 –
1. Abstract
The activity of most comets near the Sun is dominated by the sublimation of frozen
water, the most abundant ice in comets. Some comets, however, are active well beyond the
water-ice sublimation limit of ∼ 3 AU. Three bodies dominate the observational record and
modeling efforts for distantly active comets: the long-period comet C/1995 O1 Hale-Bopp
and the short-period comets (with Centaur orbits) 29P/Schwassmann Wachmann 1 and
2060 Chiron. We summarize what is known about these three objects with an emphasis on
their gaseous comae. We calculate their CN/CO and CO2/CO production rate ratios from
the literature and discuss implications. Using our own data we derive CO production rates,
Q(CO), for all three objects, in order to examine whether there is a correlation between gas
production and different orbital histories and/or size. We further examine the applicability
of existing models in explaining the systematic behavior of our small sample. We find that
orbital history does not appear to play a significant role in explaining 29P's CO production
rates. 29P outproduces Hale-Bopp at the same heliocentric distance, even though it has
been subjected to much more solar heating. Previous modeling work on such objects
predicts that 29P should have been devolatilized over a fresher comet like Hale-Bopp. This
may point to 29P having a different orbital history than current models predict, with
its current orbit acquired more recently. On the other hand, Chiron's CO measurements
are consistent with it being significantly depleted over its original state, perhaps due to
increased radiogenic heating made possible by its much larger size or its higher processing
due to orbital history. Observed spectral line profiles are consistent with the development
and sublimation of icy grains in the coma at about 5-6 AU for 29P and Hale-Bopp, and
this is probably a common feature in distantly active comets, and an important source of
volatiles for all comets within 5 AU. In contrast, the narrow CO line profiles indicate a
nuclear, and not extended, origin for CO beyond ∼ 4 AU.
– 3 –
2.
Introduction
As a typical comet approaches the inner solar system it develops a coma when frozen
water begins to sublimate at ∼ 3 AU from the Sun.1 Some comets, however, exhibit
measurable activity beyond this water ice sublimation boundary (Whitney 1955; Roemer
1962; Meech & Jewitt 1987; Hughes 1991; Sekanina et al. 1992), and up to a third of all
comets observed become active beyond 3 AU (Mazzotta Epifani et al. 2007). Sublimation
of water ice may still generate some activity beyond the normal water-ice sublimation
boundary of ∼ 3 AU. Even at very large distances, water ice sublimation can still occur
if triggered by impacts from other small bodies, tidal disruption from interactions with
massive planets, and interactions with solar flares and solar wind erosion (Strazzulla et al.
1983; Stern 1995; Sekanina et al. 1994; Boss 1994). Collectively, however, these possible
water-ice sublimation triggers are unlikely to generate the long-lived comae observed in
most distantly active comets (Mumma et al. 1993). Thus, most comets that are active
beyond ∼ 3 to 4 AU should be considered to have comae which are generated by a volatile
other than water.
Understanding how comae are generated in comets is critical to developing accurate
1This is the heliocentric distance at which the sublimation temperature of water ice,
Tsub ∼ 150 K, assuming a gas density of ∼ 10−13 cm−3, (Yamamoto 1985), equals the
blackbody temperature of a fast rotating spherical comet nucleus, Tbb, according to
Tbb = (cid:18)1 − A
ǫ (cid:19)1/4
T⊙(cid:18)R⊙
2r (cid:19)1/2
,
(1)
where A = nucleus bond albedo (typically assumed to be in the range of 0.03 – 0.15), ǫ
is the surface emissivity (∼ 0.5 - 0.9), R⊙ and T⊙ are radius and temperature of the Sun,
respectively.
– 4 –
models of nucleus composition. The activity of distant comets also has important
implications for other icy objects in the outer solar system, some of which may share
common origins with comets. For example, some short-period comets may be collision
fragments from Kuiper Belt Objects (KBOs) (Farinella & Davis 1996; Durda & Stern
2000; Schulz 2002; Dones et al. 2015). Also, KBOs, Pluto and even icy moons may also
experience similar physical and chemical conditions as distantly active comets (Weissman
1999; Bockel´ee-Morvan et al. 2001).
Analysis of visible and infrared observations often provide useful information, such as
the extended comae of distant comets are composed primarily of dust grains with a total
mass of ∼ 109kg and coma expansion velocities of 0.3 km s−1 or less (also see Whitney 1955;
Hughes 1991; Trigo-Rodr´ıguez et al. 2010; Kulyk et al. 2016). Visible lightcurves often
show variations of two to four magnitudes over a few weeks and much smaller amplitude
changes over a few hours. The larger amplitude changes are typically assumed to be due to
significant mass loss; on the other hand, smaller amplitude variations are often attributed
to the rotation state of the comet's nucleus (Whipple 1980; Meech et al. 1993).
In addition to dust, gaseous emission has also been detected. CO+, CN and OH were
the first volatiles observed in distantly active comets (Cochran et al. 1980; Wyckoff et al.
1985), but they cannot drive the activity. This is because they are produced through
ionization or dissociation of other molecules in the coma and do not exist in the nucleus.
Observations of these species, however, are still useful as tracers of their parent molecules
(Cochran & Cochran 1991; A'Hearn et al. 1984; Feldman et al. 2004). For example, spectra
of OH emission may be useful in determining whether H2O (the presumed parent of OH)
is sublimating. When CN is observed, then HCN is likely to be abundant near the surface
or in the coma. However, we cannot rule out a distributed source, such as C2N2, HC3N
or even N-bearing refractories as a parent of CN (Feldman et al. 2004; Despois et al.
– 5 –
2005; Festou et al. 1998). CO is difficult to photoionize beyond 5 AU from the Sun, so
detecting CO+ implies either unusual ionization of CO, or more likely, a very large amount
of CO being released in the coma. Thus, CO+ may be a useful tracer of CO emission
(Magnani & A'Hearn 1986; Cochran & Cochran 1991).
Carbon monoxide is of particular interest, because of its identification as a
"supervolatile" (cosmogonically abundant molecules with sufficient equilibrium vapor
pressures to vigorously sublimate at very large (> 35 AU) heliocentric distances) (see Table
1). A breakthrough in understanding distantly active comets occurred when CO gaseous
emission was detected at millimeter wavelengths in comet 29P/Schwassmann-Wachmann
1 (hereafter referred to as 29P) at ∼ 6 AU in amounts significantly higher than the
dust production rates (Senay & Jewitt 1994). Carbon monoxide was already suspected
of being in 29P's coma, because of the previously observed CO+ in 29P (Larson 1980;
Cochran & Cochran 1991) and the very low sublimation temperature of CO ice; indeed, its
detection in comets was predicted based on favorable excitation conditions and expected
production rates (Crovisier & Le Bourlot 1983). The following year, CO emission was also
detected in C/1995 O1 (Hale-Bopp) over 6.8 - 6.0 AU (Jewitt et al. 1996; Biver et al. 1996;
Womack et al. 1997), and in 2060 Chiron at 8.5 AU (Womack & Stern 1999). Doubts
were raised about the Chiron CO detection (Bockel´ee-Morvan et al. 2001); however, an
independent re-analysis (Jewitt et al. 2008) maintained that the CO line was formally
detected in Chiron. The derived CO production rates were high enough that CO outgassing
was assumed to be the main cause of activity for all three objects (Table 2).
Other highly volatile and abundant gases, such as CO2, CH4, and N2 may contribute
to distant activity, but are difficult to observe (Prialnik et al. 2004; Ootsubo et al. 2012;
Reach et al. 2013a; Bauer et al. 2015a). CO2, CH4 and N2 do not possess an electric dipole
moment and thus have no rotational transitions, which are needed for observations at
– 6 –
millimeter wavelengths. They have transitions at infrared wavelengths, but the emission
often suffers telluric contamination and are weakly excited by solar radiation beyond 4
AU. Observations from space-based telescopes indicate that CO2 is much more abundant
than CO in a large survey of distant comets (see summaries by Ootsubo et al. (2012);
McKay et al. (2012)). CO2-dominated outflows are documented in comets 103P/Hartley2
(A'Hearn et al. 2011) and 67P from spacecraft data (Hassig et al. 2015), which show that
CO2 is a major player for these comets closer to the Sun. As we discuss later, all available
data indicate that CO2 is not likely responsible for most of the observed activity in 29P,
Hale-Bopp, and Chiron. Methane has not been detected in distant comets, although
significant limits were set for 29P (see Table 2). N2 emission in comets can be estimated
from optical spectra of N+
2 (Womack et al. 1992; Lutz et al. 1993; Cochran et al. 2000;
Cochran 2002; Ivanova et al. 2016) and with mass spectroscopy from spacecraft data
(Rubin et al. 2015). All detections, or non-detection limits, point to N2 abundances being
much lower than that of CO in comets, including Hale-Bopp and 29P. In addition, the
nitrogen depletion issue in comets (Cochran et al. 2000; Iro et al. 2003) make it unlikely
that N2 plays a substantial role in driving distant comet activity.
Other volatiles have been detected beyond 3 AU, but their low production rates also
indicate that they are unlikely to generate much of the observed distant activity. For
example, CH3OH, HCN, H2CO, H2S and CS were identified in spectra of Hale-Bopp far
from the Sun, but with production rates all significantly lower than CO production rates
(Biver 1997; Womack et al. 1997).
Although most cometary studies focus on abundances from near comets, large helio-
centric coverage is also needed to aid compositional studies of all comets (Dello Russo et al.
2016). Although many measurements exist of distant comets, the causes of the activity are
not well understood. In this paper we summarize possible energy sources and observational
– 7 –
results, and make a comparative study of 29P, Hale-Bopp and Chiron. Although the dust
and gas counterparts of the comae are both addressed, the main emphasis is on the volatiles
(mainly CO), since they are directly involved in generating the distant activity. Studying
CO's behavior beyond 4 AU is especially important to confirm its natal contributions and
study it in isolation (Pierce & A'Hearn 2010).
3. Proposed Mechanisms to Generate Comae Far from the Sun
Solar radiation is the most significant energy source for comets. Consequently, the
heliocentric distance, albedo and seasonal effects of the nucleus all play vital roles in how
the energy is balanced in a comet on the surface and in the sub-surface layers of a comet
nucleus. Here we review energy sources that may play a significant role in sustained activity
at large distances, such as supervolatile sublimation, crystallization of amorphous water ice,
and radioactive decay.
First, we consider supervolatile sublimation and outgassing. If comet material started
off very cold (such as the temperatures needed initially to maintain CO, N2 or CH4 in solid
form), then their deep interiors may not have experienced any warming episodes, and highly
volatile species can be preserved in the nucleus. Sublimation of such "supervolatile" ices can
drive activity far from the Sun, and possibly beyond 40 AU for the most volatile CO. One
concern raised about supervolatile ices is that severe devolatilization may have occurred
due to significant heating over the ∼ 4.6 Gyr since the comets formed. As Kouchi & Sirono
(2001) point out, however, laboratory results indicate an extremely low thermal conductivity
for water ice in comets, suggesting that heating is negligible below the outer several tens of
centimeters, and so extends the lifetime of supervolatiles in comet nuclei. Additionally, we
point out the abundant evidence for cometary supervolatiles, as recorded by observations
and space mission experiements, which show that supervolatiles are long-lived in cometary
– 8 –
nuclei and have not been exhausted close to the Sun or preferentially in comets that sustain
more heating (such as Jupiter Family comets, which have low orbital inclinations and
orbital periods shorter than ∼ 20 years) as compared to new comets (A'Hearn et al. 2011,
2012; Hassig et al. 2015; Dello Russo et al. 2016).
Next, we examine the amorphous water ice crystallization phase change in comet
nuclei. If cometary water originates primarily from preserved interstellar grains, or from the
distant outer regions of the solar disk, then it is likely to accumulate in an amorphous state
(Kouchi et al. 1994; Irvine et al. 2000). Upon heating, such ice undergoes a solid phase
transition to a crystalline structure. The lab-measured threshold for transition to cubic ice
is ∼100 K, with another transition from cubic to hexagonal ice at ∼160 K (Laufer et al.
1987). However, only the first transition is exothermic and it involves a significant change
in specific density and arrangement of ice structure, so that trapped molecules can be
released. The functional form for the amorphous-crystalline transition, with 95% of the ice
changing to the crystalline phase in a time (sec) tc, was derived as
tc = 9.54 × 10−14 exp(5370/T )
(3)
at a temperature T (K), for the range of 125-150 K, (Schmitt et al. 1989). Extrapolation
to lower temperatures gives 120 K as the activation temperature, where the rate becomes
rapid (∼1 month). This is commonly used as a nominal crystallization temperature.
This phase transition could release enough energy, approximately 109 ergs g−1
(Ghormley 1968; Klinger 1981), to increase the temperature of the surrounding ice by
another 45 K (Enzian et al. 1997). This, in turn, could trigger even more crystallization
and sublimation, creating a runaway crystallization effect (Patashnick 1974; Prialnik et al.
1992). Accordingly, much of the outermost ice of a comet that has been heated to ∼ 140
K is expected to be in the crystalline form. This could mean that most, if not all, of the
– 9 –
comets that make repeated approaches within 5 AU 2 of the Sun should have crystalline
water ice on the outermost layers of the nucleus.
In addition to releasing energy, this phase transition decreases the density of ice from
ρa ∼ 1.2 g cm−3 to ρc ∼ 0.9 g cm−3. This structural change releases gas molecules that were
trapped in the original ice, such as CO and CO2 (Bar-Nun et al. 1988; Bar-Nun & Owen
1998; Prialnik et al. 1995).
Runaway amorphous ice crystallization is modulated by the fact that laboratory studies
show that exothermic crystallization occurs only for pure amorphous ice. Slight impurity
levels (such as > 2% in mol of CO in the ice matrix) in the ice are enough to reduce the
effective latent heat of crystallization, thus reducing the exothermic nature of the process
(Kouchi & Sirono 2001). This could result in isolated bursts, rather than crystallizing all at
once. Crystallization of amorphous water ice is a mechanism included in many theoretical
models of distantly active comets (e.g., Prialnik & Bar-Nun 1990a,b; Capria et al. 2000;
Sarid et al. 2005). While a few other exothermic mechanisms have been suggested, such
as polymerization of condensed-phase HCN (e.g., Rettig et al. 1992; Matthews 1995)
and melting and dissolution of gases (Miles 2016), these processes have not been put to
a quantitative test or included in a full account of the energetics involved in cometary
evolution.
Radiogenic heating is an additional source of energy in distant comets that may deplete
significant amounts of frozen CO and other supervolatiles from the nucleus (Prialnik et al.
1987; Haruyama et al. 1993; Prialnik et al. 1995; De Sanctis et al. 2001a). The effectiveness
of this mechanism depends critically on such parameters as the diffusivity and thermal
2Four to five AU is the approximate heliocentric range where a nucleus is expected to
reach 140K due to solar heating, see equation 1.
– 10 –
conductivity of the nucleus, both of which have a large range in values for the ice and silicate
components. However, if the nucleus is smaller than a radius of ∼10 km, then the cooling
timescale from radiogenic heating is shorter than the heating timescale. These objects
would effectively get rid of the excess energy generated by radioactive decay quicker than
this energy can promote a thermodynamic heating cycle on icy layers (Prialnik & Podolak
1995; Prialnik et al. 2008). Hence, for comets with nuclei smaller than 10km, the initial
temperature is probably preserved deeper than about 100 meters.
One important concern with modeling cometary ice is that the experimental value of
the thermal conductivity of water ice is 104 times lower than the theoretical value (Klinger
1980; Kouchi et al. 1992), yet both quantities have been used in models of distantly active
comets. The grain packing configuration and appreciable porosity can change the thermal
conductivity by orders of magnitude. It is affected by the porosity, Hertz factor (the contact
area relative to cross-sectional area of material grains) and the specific thermal conductivity
of the ice (what phase of ice and its level of impurity) and pores (depends on the shape
and inter-connectivity of the porous space). The surface-to-volume ratio of the pores is also
sensitive to the actual pore size distribution, at any given layer of the nucleus, and can vary
by a few orders of magnitude (Sarid et al. 2005).
If CO sublimation generates activity in a distant comet, a simple estimate of the
sublimating area, a, on the nucleus can be derived using the following equation, assuming
that the nucleus is a rapid rotator
a =
(dM/dt)r2LCO
F⊙(1 − A)
,
(4)
where dM/dt is the total mass loss rate of CO, F⊙ = 1360 W/m2, r is heliocentric distance
in AU, LCO is the latent heat of sublimation for CO, and A is the Bond albedo of the
nucleus. For example, using dM/dt = 2000 kg s−1 from the CO observations, this estimate
gives a circular spot with an emitting radius of ∼ 2 - 6 km, depending on whether the CO
– 11 –
is released as a sublimating ice (LCO ∼ 4 × 105 J/kg), or as a trapped gas in water ice
(LCO ∼ 1.5 × 106 J/kg) (Brown & Ziegler 1979), corresponding to < 0.1% of the surface
area for 29P and Hale-Bopp.
As a cautionary statement, one should consider that sometimes volatiles detected in
the coma did not originate directly from the comet's nucleus, and thus do not play a role in
generating activity. For example, a source could be ice-coated grains of refractory material
expelled from the nucleus (Delsemme & Miller 1971; Hanner 1981, 1984). Once in the solar
radiation field, there is enough energy available (if within ∼ 5-6 AU of the Sun) for the
ice-coatings to sublimate. Such grains reportedly are destroyed very quickly (within 20
hours for 1 micron sized grains) once their temperatures reach ∼ 150K (Lamy 1974).
Evidence for the sublimation of icy grains was seen in the emission of OH at 308 nm
in comet Bowell, which was detected as the comet moved from 5 to 3 AU (A'Hearn et al.
1984, 1995). Icy grains with radii of 7 - 30 micron were determined to exist in Hale-Bopp's
coma with a median lifetime of about two days at 2.9 AU (Lellouch et al. 1998). Further
evidence is found from spectral line-widths of methanol and other species in Hale-Bopp at
∼ 5 AU, which were much wider than one would expect for molecules sublimating directly
from the nucleus (Womack et al. 1997). 103P/Hartley 2 was also observed by the EPOXI
mission and the large effective sublimation area derived from photometric observations was
determined to be from ice grains sublimating, but at much closer distances to the Sun
(Protopapa et al. 2014; Kelley et al. 2013).
4. Observations of Distantly Active Comets
In this section, we review what is known about each object, and in the next section
we examine the similarities and differences between these objects and propose a common
– 12 –
model that may explain the distant activity.
4.1. Comet Hale-Bopp
Comet Hale-Bopp presented an exceptional opportunity to study distant activity,
since it already had a giant coma upon discovery at 7 AU. It is a long-period comet with
a high-inclination orbit, and, thus has a very different orbital history than either 29P or
Chiron. Because it has spent most of its lifetime very far from the Sun, it is probably
less processed than many other comets. It is a very large comet - its nucleus radius was
measured to be in the range of R = 20 - 35 km (Weaver & Lamy 1997; Sekanina 1997;
Fern´andez et al. 1999), and RHale−Bopp = 30 ± 10 km was recommended as a compromise
(Fern´andez 2000), which we adopt for calculations in a later section. Hale-Bopp's radius is
very similar to 29P, but much smaller than Chiron.
Hale-Bopp exhibited several changes as it approached the Sun. There were reports
of a swirled "jet" and a few outbursts in the dust coma from 7 to 6 AU (Sekanina
1996; Weaver et al. 1997). Multiple porcupine-like features were first observed during
May-Jul 1996 at r ∼ 4 AU and continued for several months during perihelion approach
(Boehnhardt et al. 1997; Braunstein et al. 1997; Weaver et al. 1997). Surface brightness
profiles of Hale-Bopp's coma at ∼ 7 AU were fit with a power law and had a slope of s
= −0.9 to −1.2 (Weaver et al. 1997), which is consistent with a spherically symmetric
outflowing coma subject to solar radiation pressure.
Numerous records were set for the detections of cometary gases at large heliocentric
distances with Hale-Bopp, including the first time that CH3OH, HCN, H2S, CO2 and
H2CO were observed in a comet beyond 4 AU (Womack et al. 1997; Biver et al. 1997b;
Woodney et al. 1997; Crovisier et al. 1997). Many of these species were observed at even
– 13 –
larger distances post-perihelion (Biver et al. 1997b, 2002). Carbon monoxide emission
was first detected in Hale-Bopp at ∼ 6.8 AU (Jewitt et al. 1996; Biver et al. 1996) with
production rates high enough to cause the observed dust production. CO2 emission was
detected at 4-5 AU via its vibrational bands at 4.3 µm and inferred from CO Cameron band
emission, present at ∼ 15% of the production rate of CO (Crovisier 1997; Weaver et al.
1997). A strong upper limit was obtained for N+
2 in Hale-Bopp near perihelion by
Cochran et al. (2000), which when compared with CO+ measurements is also consistent
with N2 produced in much smaller quantities than CO. The production rates of OH (a
proxy for H2O), CH3OH, HCN, H2S and H2CO were all much lower than those of CO. We
conclude that Hale-Bopp's gas coma was dominated by CO whenever the comet was beyond
4 AU pre-perihelion (Weaver et al. 1997; Womack et al. 1997; Biver et al. 2002).
The physical state of Hale-Bopp's coma volatiles is revealed by their millimeter-
wavelength spectral line profiles. For example, CO spectra had a velocity blue-shift of ∼ 0.4
km s−1, and narrow line-widths of FWHM ∼ 0.3-0.4 km s−1 (Biver et al. 1996; Jewitt et al.
1996). The narrow line-widths are consistent with originating from deeper and colder
regions in the nucleus. At larger distances, the CO line profile occasionally exhibited two
narrow blue-shifted peaks (Womack et al. 1997; Gunnarsson et al. 2003); however, a single
peak was observed most of the time. The velocity vector component of the spectra shows
a sunward release of CO into the coma. The CO line-shape was different post-perihelion
and exhibited a red-shifted peak along with a blue-shifted peak. The double peaks were
interpreted by Gunnarsson et al. (2003) as evidence for both nuclear and extended CO
sources, similar to their model for 29P.
Interestingly, beyond ∼ 4 AU, the spectral line profiles of most other species were
significantly different from that of CO. For example, CH3OH, OH, HCN and H2S had
much broader line-widths than CO, and very little, if any, velocity shifts (see Figure 1).
– 14 –
Their line-widths were FWHM ∼ 0.7 to 1.5 km s−1 over the range of 5 to 4 AU. Beyond
4 AU, CO linewidths were typically much narrower, 0.4 km s−1, although occasionally
the line was observed to be much broader (Womack et al. 1997; Biver et al. 2002). Since
the spectral linewidths are mostly a result of the thermal distribution of the gas, different
production mechanisms between CO and the other molecules are responsible for most of
the observed variation in line profiles. The broader, cometocentric line profiles may be
explained if H2O, CH3OH, HCN, H2CO and H2S were first released into the coma on
grains, and then sublimated when they accumulated enough energy from the solar radiation
field. Volatiles released from a symmetric halo surrounding the nucleus will appear centered
on the ephemeris velocity. They will also be at a higher temperature and have a larger
velocity distribution (and larger line-widths) than those released directly off the nucleus.
Occasionally, HCN and CH3OH spectra mimicked the CO line profile shape, which may
indicate that they were produced from two or more methods (Gunnarsson et al. 2008).
Unfortunately, the CO2 band at 4.3 microns in Hale-Bopp is not spectrally resolved and,
so, one cannot derive a useful upper limit to velocity information of the CO2 emission.
The pre-perihelion production rates and narrow line profiles of CO, and dust production
rates, in Hale-Bopp are best reproduced by nucleus models that contain amorphous water
ice with trapped CO gas, small amounts of CO ice, and crystallization of water ice starting
at ∼ 7 AU from the Sun (Prialnik & Podolak 1999; Enzian 1999; Capria et al. 2002). Prior
to crystallization, sublimation of solid CO is invoked as the dominant source of activity,
without detailing the source of CO ice in the nucleus. The models predict slightly higher
production rates post-perihelion than pre-perihelion, which was not seen by Biver et al.
(2002); Gunnarsson et al. (2003). This suggests that thermal inertia was not very important
for Hale-Bopp once the comet was beyond 4 AU, and therefore most of the activity is
at or near the surface of the nucleus. Observations are also consistent with the activity
being well-correlated with insolation, as evidenced by symmetric behavior of gas and dust
– 15 –
production rates before and after perihelion (Biver et al. 2002).
4.2. Chiron
2060 Chiron is an unusual icy minor planet. When it was discovered in 1977, it was
categorized as the most distant known asteroid. Its reflectance spectrum had a relatively
neutral "color", similar to C-class asteroids (Lebofsky et al. 1984; Hartmann et al. 1990;
Stern et al. 2014), and it occasionally outbursts and develops a coma, which warrants it a
formal comet designation (Tholen et al. 1988; Meech & Belton 1990). Evidence for CN and
CO gaseous emission (Bus et al. 1991; Womack & Stern 1999) further secured the cometary
identity.
In addition to its cometary designation, orbital dynamical studies place Chiron
among the small population of large, outer solar system objects called Centaurs, which
are moderately sized Kuiper Belt Objects that have moved in closer, with unstable
orbits, and cross one or more giant planet orbits (e.g., Fernandez 1980; Stern 1989;
Hahn & Bailey 1990; Levison & Duncan 1994; Fernandez & Gallardo 1994). Its brightness
varies significantly on both short and long timescales, with changes in V and R mag of ∼
0.05 − 0.10 magnitudes over hours and > 1.5 mag over months and years (Tholen et al.
1988; Bus et al. 1989; Luu & Jewitt 1990; Buratti & Dunbar 1991; Duffard et al. 2002). A
water-ice signature was detected in the near-infrared, which also appears to vary over time
(Foster et al. 1999; Luu et al. 2000; Romon-Martin et al. 2003). It is not known whether
the ice is primarily crystalline or amorphous; however, the measured color temperature
(the equivalent blackbody temperature that is derived from the continuum spectrum of
an object) of the nucleus is high enough (125 - 140 K) to trigger the crystallization phase
change (Campins et al. 1994).
– 16 –
One of the most puzzling aspects of Chiron is that none of these changes are correlated
well with heliocentric distance. In fact, Chiron reached peak brightness in the visible in 1989
at ∼ 11 AU, several years before perihelion, and it may have been even brighter than this
at aphelion, 19 AU (Bus et al. 1993). For decades, the light curve was largely unexplained
- other than some of the short-term variations being attributed to reflection off a rotating
nucleus (Bus et al. 1989; Marcialis & Buratti 1993; Lazzaro et al. 1997). Attempts were
made to fit the remaining light curve contributions with a model of amorphous water ice
and a highly volatile ice, such as CO and/or CH4, with a dust mantle (Fanale & Salvail
1997), but they could not explain the lack of correlation with heliocentric distance.
Recently, however, these unusual changes in light curve and spectra have been explained
as due to reflected light from a ring system circling Chiron, based on occultation data
(Ortiz et al. 2015; Ruprecht et al. 2015). The proposed rings can account for most of the
short- and long-term brightness variations, largely by a change in their orbital tilt, if the
ring particles have a different albedo than the surface. This may also explain the observed
changes in the water-ice spectra, provided the ice is in the rings and not on the surface.
Interestingly, another large Centaur, Chariklo, has a ring system, and similar changes are
noted there (Duffard et al. 2014). If confirmed in Chiron, this suggests a paradigm shift,
where rings may be commonly found around Centaurs and perhaps other large icy objects,
like Kuiper Belt Objects, and rings need to be included as possible contributors to observed
differences among KBOs. We note, however, that the Pluto system does not have any rings
and the New Horizons project searched very thoroughly for them. So, while rings could be
contributors, for some of our best examples of KBOs, rings are not present.
The evidence is very strong for a model where most of Chiron's brightness changes are
due to the changing orbital tilt of a ring system and nucleus rotation; however, these cannot
explain the occasional outbursts and development of a coma. One possibility is that the
– 17 –
rings and coma may even be interdependent, as the most likely formation mechanism for the
rings is proposed to be a debris disk fed by some cometary activity, with a possibility that
in-falling material from the rings could trigger outbursts (Ruprecht et al. 2015). If instead
of rings, Ruprecht et al. (2015) also raise the possibility that narrow jets may generate
activity by a low-gravity analog to the geyser-like vents that are sometimes observed on
Neptune's large satellite Triton or on Pluto, fueled by sub-surface geological activity. More
observations are needed of Chiron, in and out of outburst, to confirm the rings or jets.
Chiron's outbursts and coma occur at distances from the Sun too far to be caused by
H2O and CO2 ice sublimation, the drivers of most normal comets near the Sun. Instead, the
outgassing activity is probably largely due to the sublimation of even more volatile species.
CO is the most likely candidate. Its emission in Chiron was singularly detected via the 1-0
rotational line in 1995 with production rates high enough to drive the observed dust coma
(Womack & Stern 1999; Jewitt et al. 2008). The measured CO production rate indicates
that much less than 1% of Chiron's surface was active in 1995, consistent with CO-driven
vents or geysers.
Occultation and thermal emission measurements indicated that Chiron has a very
large diameter, ranging from D = 150 – 220 km (Campins et al. 1994; Elliot et al. 1995;
Bus et al. 1996; Fern´andez 2000). Following the discussions by Groussin et al. (2004) and
Fornasier et al. (2013), we recommend a diameter of DChiron=218±20 km, which we use for
later calculations. Chiron's nucleus is clearly larger than those of Hale-Bopp and 29P, and
significantly larger than most comet nuclei.
Chiron has a low, but steady dust production rate of ∼ 0.5 kg s−1, with a mean
speed of ∼ 100 m s−1 for a coma that is populated by 1-micron sized grains, according to
analysis of near-infrared H- and R-band observations (Luu & Jewitt 1990). On the other
hand, modeling of V-band images indicated that the coma was dominated by much larger
– 18 –
100-micron sized dust grains that were largely un-sampled by the near-infrared data (West
1991; Fulle 1994; Romon-Martin et al. 2003). The larger grain sizes lead to a much higher
mass loss rate of 20 ± 10 kg s−1, and slower dust expansion velocities of ∼ 5 to 10 m s−1.
This is a very low dust production rate when compared to 29P or Hale-Bopp.
Interestingly, Chiron's color temperature measurements are 40 - 50 K warmer than
expected for its nucleus if one assumes an equilibrium temperature for a fast rotator
spherical blackbody. It is not known whether the aforementioned ice is primarily crystalline
or amorphous; however, the measured color temperature of the nucleus is high enough
(125 - 140 K) to trigger the crystallization phase change (Campins et al. 1994). Unlike
29P, Chiron's nucleus rotation period is well determined to be 5.917813 ± 0.000007 hours
(Marcialis & Buratti 1993).
Information about Chiron's gas coma is scarce. Thus far, CN and CO are the
only species seen in the coma, and both were detected only once (Bus et al. 1993;
Womack & Stern 1999). The optical CN detection of ∼ 4 × 1025 mol s−1 at 11 AU correlates
to a high production rate for a comet at 11 AU, and >4 times greater than Hale-Bopp at
10-12 AU (Rauer et al. 2003). Emission from CN was searched for when Chiron was near
perihelion at ∼ 8.5 AU, and was not detected down to a limit that is ∼ 25% lower than
what was observed at 11 AU (Rauer et al. 1997). Chiron was reported to be undergoing
an outburst during both observing runs, so it is puzzling that CN was detected only once.
Thus, the CN production in Chiron may be variable and not well-correlated with either the
dust production rate or heliocentric distance. The millimeter-wavelength CO detection in
Chiron is consistent with a production rate of ∼ 1028 mol s−1 at 8.5 AU (Womack & Stern
1999), which is also similar to Hale-Bopp at the same distance (Biver et al. 2002), see
Figure 2. The observed CO line-width was very narrow and similar to what has been
observed for 29P and Hale-Bopp. The CO line-width was FWHM = 0.39 km s−1; however,
– 19 –
the line was not resolved, and this should be considered an upper limit to the true spectral
line-width. The narrow line width is indicative of emission from a cool gas, T < 110 K, and
is also consistent with a dusty CO coma, in which dust heats gas in the near-nucleus region,
resulting in terminal gas velocities of ∼ 0.4 km s−1 (Boice & Huebner 1993).
An upper limit of Q(CO) < 2 × 1027 mol s−1 was obtained over five observing runs
from Mar 1998 - Jul 2000, when Chiron was 9 - 11 AU (Bockel´ee-Morvan et al. 2001). This
limit is almost ten times lower than the production rate measured several years earlier by
Womack & Stern (1999). Hence, the CO detected in 1995 may have been the result of an
outburst and not caused by any long-lived activity from Chiron, a conclusion also reached
by Bockel´ee-Morvan et al. (2001).
A detailed study by Fornasier et al. (2013) determined that a gas production rate of
at least 1-10 × 1027 s−1 were needed to drive dust production rates derived from optical
observations of the coma. This amount is consistent with the CO production rate measured
at 8.5 AU and comparable to upper limits derived from ∼ 9=11 AU. Fornasier et al. (2013)
also raise CO2 as capable of driving the observed dust activity. Unfortunately, there are no
detections or upper limits to CO2 production rates in Chiron.
The high color temperature of Chiron's nucleus and the CO and dust production rates
are predicted by two models with a highly porous nucleus containing dust and gas-laden
amorphous water ice with more volatile ices deeply buried (Prialnik et al. 1995; Capria et al.
2000). In these models, crystallization of amorphous ice in the nucleus is proposed to
occur near the surface, which then releases the trapped CO gas. This may be the preferred
mechanism, since at the surface temperature of Chiron (120K, see Campins et al. 1994) CO
and CO2 ices would not survive shallower than a few skin depths (1 skin depth is ∼20 m,
assuming a density of 0.7 g/cm3, see Prialnik et al. 2008). Thus, such solid ice phases of
these volatile species can be ruled out as significant constituents of the top surface layers,
– 20 –
and are probably restricted to deeper sections of the nucleus.
4.3. Comet 29P/Schwassmann-Wachmann 1
Comet 29P/Schwassmann-Wachmann 1's coma has been documented for 90 years. The
comet appears to be nearly always active with numerous reports of dramatic outbursts of 4
to 8 magnitudes over relatively short time periods (e.g., Roemer 1958, 1962; Whipple 1980;
Hughes 1991; Trigo-Rodr´ıguez et al. 2010). Originally classified as a short-period comet, it
has a low inclination orbit, and its heliocentric distance ranges between 5.7 < r < 6.3 AU.
Its current orbit is nearly circular (e = 0.04) and is relatively near the water ice sublimation
zone. Dynamical studies show that 29P's orbit has become more circular over time, and
it probably only recently transferred to this orbit from one much farther from the Sun,
possibly the Kuiper Belt. It is often considered to have a Centaur orbit (e.g., Fernandez
1980; Stern 1989; Levison & Duncan 1994; Fernandez & Gallardo 1994).
A recent study of 29P's nucleus radius derives R29P =30.2+3.7
−2.9 km (Schambeau et al.
2015), which we use for later calculations in this paper. This puts 29P at about the same
size as Hale-Bopp and much smaller than Chiron. 29P's large asymmetrical dust coma
often exceeds 400,000 km across, and spiral structures often seen in the coma are consistent
with one or more active areas on the nucleus (Roemer 1958; Whipple 1980; Cochran et al.
1982; Reach et al. 2013a). Optical imaging techniques implied that the dust coma had a
surface brightness profile (see Eq. 1) of s = −1 to −1.5 (Meech et al. 1993; Jewitt 1993),
consistent with a steady-state, uniformly outgassing coma subject to some solar radiation
pressure. Jewitt (1990) estimated a dust mass loss rate of 10 kg s−1 and dust expansion
velocity of 200 m s−1 from images. However, this may be an underestimate of the total
mass loss, since the reflected light is dominated by light scattered by grains with sizes
comparable to the observed wavelength. As a result, this technique may undercount larger
– 21 –
grains, which probably play an important role in the comet's mass loss. One model of the
grain size distributions indicates a much higher mass loss rate of ∼ 600 kg s−1 (Fulle 1992).
CN is often the first emission feature detected in optical spectra of comets far from the
Sun, so it was a surprise when CO+ was the first gaseous species seen in 29P (Cochran et al.
1980). CO+ emission in 29P was initially believed to be explained by photoionization
of CO; however, the lifetime for this process at 29P's distance from the Sun (Larson
1980; Cochran & Cochran 1991) is too long to account for the observed column densities.3
Collisional ionization of CO from high energy solar wind electrons has been proposed,
instead, to explain formation of the CO+ in 29P (see Cochran & Cochran 1991). CN was
eventually detected in 29P in Dec 1989 with a line strength weaker than the neighboring
CO+ lines (Cochran & Cochran 1991; Cook et al. 2005).
The rotational period of 29P's nucleus is unknown, with measurements ranging from 14
hours to longer than two months (Miles 2016; Ivanova et al. 2015; Moreno 2009; Meech et al.
1993; Stansberry et al. 2004; Jewitt 1990; Whipple 1980). The large uncertainty in 29P's
rotation period is a significant problem for constraining outgassing models. Since 29P's
orbit has a very low eccentricity, the surface experiences a nearly constant rate of heat
forcing (and cooling rate). Thus, the variations in illumination for any given patch on
the surface are predominantly a function of period and obliquity. The co-latitude and
co-longitude dependence drives the sub-surface activity (neglecting lateral heat conduction).
The uniqueness of 29P's activity pattern may be the result of its relatively circular orbit
3The lifetime against photoionization for CO at 29P's distance ranges from 3 - 8 x107
sec, depending on the solar activity (Fox & Black 1989). At the average rate of 5 × 107 sec,
CO molecules moving at 0.45 km s−1 away from the nucleus would travel ∼ 0.1 AU from the
comet before being photoionized, which is well beyond the region where CO+ is observed
(Festou et al. 2001).
– 22 –
and a non-simple spin state (possibly a very slow rotator with a highly oblique spin pole).
The first direct evidence for a volatile that could generate a coma in 29P was the
detection of CO gas (Senay & Jewitt 1994). A production rate of Q(CO) = 3.0 x 1028
molecules s−1 was derived, which corresponds to a mass loss rate of ∼ 2000 kg s−1. Since
this is much greater than the dust mass loss rate of 10 - 600 kg s−1, it is assumed that CO
is driving the observed activity. Upper limits of Q(CH4) < 1.3x1027 mol s−1 (< 5% of CO)
were determined from infrared spectra of 29P (Paganini et al. 2013).
Similarly, sensitive upper limits to CO2 emission show that this species is not dominant
in 29P's coma with Q(CO2) < 80 Q(CO) (Ootsubo et al. 2012; Reach et al. 2013a;
Bauer et al. 2015a). We conclude that N2, CH4 and CO2 are produced in minor amounts
when compared to CO. Comet 29P seemingly always has at least a minimal CO production,
of Q(CO) ∼ 3 × 1028 mol s−1, which is often referred to as the quiescent or non-outbursting
state (Festou et al. 2001).
Sporadic outbursts also occur which increase the comet's CO production by up to a
factor of seven (see Figure 2), even though the comet's orbital distance does not change
very much during this time. Clearly, CO outgassing is tied to the outbursting mechanism
in 29P. Surprisingly, no correlations were found between observed CO+ emission and the
outbursting state, as evidenced by optical observations of the coma. The production
mechanism for CO+ in this comet is not clear (Cochran et al. 1980, 1982). Simultaneous
observations of CO and CO+ are needed to better understand coma chemistry.
The CO activity in 29P was proposed to be released from a relatively small active area
on the surface by Senay & Jewitt (1994), either by sublimation of CO, or as trapped in
water ice. However, detailed models of frozen CO sublimating (whether pure or adsorbed
onto water ice) does not reproduce the observations well (i.e. Enzian et al. 1997) and
the surface is too warm for CO ice to survive for very long. Different models exist that
– 23 –
incorporate energy provided by HCN polymerization and the crystallization of amorphous
water ice (Gronkowski & Smela 1998) or melting and dissolution of gases (Miles 2016);
however, they do not include a full account of the energetics, so these models are of
limited value. Moreover, the dust outflow velocities predicted are approximately three
times lower than what was measured from HST WFPC2 imaging data (Feldman et al.
1996). The model which best explains the CO production rates to date is one that includes
crystallization of amorphous water ice with trapped CO gas and pockets of CO ice, a highly
tilted rotation axis and surface erosion (Enzian et al. 1997). In this model, CO production
primarily comes from release as a trapped gas when crystallization occurs in amorphous ice
in the sub-solar area. This model also explains outbursts with surface erosion where dust
grains are released from the ice-dust matrix by the escaping gas.
Further clues about the physical state of the coma are given by the millimeter and
infrared spectra of CO. The CO spectral line shapes in 29P (in both the quiescent and
outbursting states) are frequently characterized by two very narrow (FWHM line-width
∼ 0.10 to 0.15 km s−1) velocity components, which are found at ∼ +0.5 km s−1 and -0.3
km s−1 with respect to the cometocentric velocity (Crovisier et al. 1995; Biver et al. 1997a;
Festou et al. 2001; Gunnarsson et al. 2002, see Figure 3). Two emitting regions have been
proposed to explain the double-peaked (red and blue-shifted) CO line profile: a day-side
and night-side region (Crovisier et al. 1995). In contrast, the two peaks are interpreted
by (Crifo et al. 1999) as due to temperature distributions on the surface, and not the
spatial location of vents. Still another idea is that the line-shape may be partly due to
under-sampling the CO emitting region, which would preferentially detect the gas with the
greatest Doppler shifts (Festou et al. 2001). Finally, the two lines could be caused by two
different sources of CO: a nuclear and an extended source originating from icy grains in the
coma (Gunnarsson et al. 2002).
– 24 –
Raster (point-by-point) mapping data with the SEST 15-m telescope indicated that
there was a substantial secondary source of CO in the coma, which was hypothesized as
being due to the release of CO from icy grains in the coma in 29P (Festou et al. 2001;
Gunnarsson et al. 2002). In this scenario, the sunward emission from the nucleus gives a
strong blue-shifted peak, and the isotropic emission of the extended source gives a so-called
"horned" appearance due to under-sampling of the icy grain emission source by the radio
beam. The blue-shifted part of the "horn" is blended with the sunward peak. With this
model, the extended source produces ∼ 3 times the amount of CO as the nuclear source
and is the dominant source of CO emission.
Subsequent mapping of CO emission with the IRAM 30-m dish did not show evidence
for such a widespread secondary source (Gunnarsson et al. 2008). However, the data
showed an asymmetric localized feature in the coma with "excess emission. This is evident
in both the dust and gas comae and extends 0.5 to 2 arcminutes south of the nucleus
(Schambeau et al. 2015; Gunnarsson et al. 2008), which corresponds to a projected distance
starting at ∼ 1.1 × 105 km from the nucleus (we find that the published number in
Gunnarsson et al. (2008) is too high by a factor of ten). This is sometimes also referred to
as a jet feature. Thus, it is possible that 29P split off a smaller piece or fragment during 2003
- 2004. Interestingly, CO emission was detected from both the main and fragment sources
by Gunnarsson et al. (2008). The fragment was found to have a slightly smaller blueshift
(-0.18 km/s) than the main nucleus component (-0.35 km/s). The spectral line profile
of the secondary also showed the classic horned appearance, consistent with the spectra
obtained at the primary position. Follow-up high angular resolution spectral mapping, or
interferometry, would be invaluable for studying this "excess emission" phenomenon in 29P.
The true nature of this enhanced CO emission region in the coma is not known. A
long lasting, but largely hidden, fragment or secondary component could be an important
– 25 –
additional source of CO in 29P's coma (Gunnarsson et al. 2008). A secondary component
would also complicate efforts to measure the rotation period of the main nucleus, which has
so far remained undetermined.
We note that 174P/Echeclus, another Centaur that is sometimes active, reportedly
had a distinct secondary source, which may have recently split from the primary nucleus
(Rousselot 2008; Fern´andez 2009). This secondary extended emission in Echeclus was at
a similar distance from the nucleus: the angular separation is a factor of ∼ 3 times, and
the derived spatial separation is ∼ 1.5 times that of 29P's excess emission (calculated
from Rousselot 2008; Gunnarsson et al. 2008) using the correct distance (see previous
paragraph). Processes acting at 29P may be similar to those at Chiron, the other Centaur
in this study, producing secondaries and rings.
Infrared and millimeter wavelength spectra also showed that the CO inner coma gas in
29P is at an extraordinarily low rotational temperature of ∼ 5K (Gunnarsson et al. 2008;
Paganini et al. 2013), making it one of the coldest gases known in the the solar system.
As discussed in Gunnarsson et al. (2002); Paganini et al. (2013), if CO molecules start out
from the nucleus with Trot ∼ 5K, they will change excitation conditions over time and
reach fluorescence equilibrium within ∼ 2 days. Therefore, many millimeter wavelength
observations with larger beamwidths are recording spectra from molecules which are close
to fluorescence equilibrium with a higher rotational temperature of 10-15K.
29P's coma is predicted to have a significant amount of water due to the sublimation
of icy grains, if water ice is present at the surface and if grains are ejected (Crifo et al.
1999). Initially, searches showed no water down to a limit of Q(OH) < 2-3 × 1028 mol
s−1 (Cochran & Cochran 1991; Feldman et al. 1996). 29P was observed with the Herschel
Space Observatory, and preliminary results claim a detection of water and HCN with line
profiles that are consistent with being emitted from icy grains in the coma, but no values
– 26 –
have been published yet for the production rates (Bockelee-Morvan et al. 2010, 2014).
5. Discussion and Comparison of Distantly Active Comets
In this section, we review the similarities and differences in the observational data
of Hale-Bopp, 29P and Chiron, and discuss the models that best fit these data. We
focus on the spectroscopic results, mainly production rates and line profile characteristics.
Representative production rates of important species in 29P, Chiron and Hale-Bopp are
summarized in Table 2. In particular, we 1) examine the production rate ratios of CN/CO
(and HCN/CO) and 2) CO2/CO, 3) discuss clues from spectral line profiles of volatile
species, and 4) recalculate CO production rates for all 3 comets with consistent modeling
assumptions. We then 5) summarize observational constraints to models of distant activity
and 6) analyze Q(CO) for all three objects when adjusted for nucleus size and heliocentric
distance.
5.1. Production Rate Ratios: CN/CO and HCN/CO
CN and CO are the only gaseous species observed in all three comets, and we briefly
compare their relative production rates to test models of CN and HCN in comae and
to assess whether HCN is a likely competitor for triggering activity in these objects.
Using values from Jewitt et al. (1996); Fitzsimmons & Cartwright (1996); Rauer et al.
(1997); Biver et al. (2002), we calculate that the ratio Q(CN)/Q(CO) was ∼ 0.003 in
Hale-Bopp at ∼ 6.8 AU pre-perihelion, and the same value at 4.6 AU post-perihelion
(see Table 3). We derive the ratio Q(HCN)/Q(CO) ∼ 0.004 at 6 AU post-perihelion
using values from Rauer et al. (1997); Biver et al. (2002). Interestingly, Q(HCN)/Q(CO)
∼ Q(CN)/Q(CO), even at somewhat different heliocentric distances. This is consistent
– 27 –
with work by Rauer et al. (2003); Fray et al. (2005); Bockel´ee-Morvan et al. (2008), which
conclude that Hale-Bopp's CN largely originated from HCN photolysis. Additionally, it is
clear that HCN was produced in much smaller amounts than CO in Hale-Bopp at large
heliocentric distances, and thus does not compete with CO in triggering activity.
Unfortunately, it is not clear what the CN/CO production rate ratio is for Chiron,
because CN and CO were not measured at the same time. However, if we use the CN value
at 11 AU and CO measurement at 8.5 AU, we get a similarly low value of Q(CN)/Q(CO) ∼
0.003. However, we stress that these were not simultaneous measurements for Chiron, and
hence the ratio is probably of limited value.
CN is only rarely observed in 29P and, when seen, it appears depleted with respect
to CO, when compared to Hale-Bopp, which we use as a proxy for an unprocessed new
comet. Using the CN production rate from Cochran & Cochran (1991) and the average
non-outbursting value of Q(CO) ∼ 3 × 1028 for 29P, we calculate a production rate
ratio Q(CN)/Q(CO) ∼ 0.0003, which is almost ten times lower than what is derived
for Hale-Bopp at 4 and 6 AU. By this metric, 29P appears to have much less CN (and
presumably HCN) than CO when compared to Hale-Bopp and possibly also Chiron.
As mentioned previously, 29P was observed with the Herschel Space Observatory, and
preliminary results claim detections of H2O and HCN (Bockelee-Morvan et al. 2010, 2014).
Once HCN production rates are published from these Herschel data, they will be useful for
comparing with measured CO and CN production rates to test models of release of all three
species from the nucleus (McKay et al. 2012). It will be also interesting to see whether the
Herschel data support CN arising from HCN in 29P. The relative amounts of HCN and CO
produced from 29P could provide clues to the comet's formation environment. For example,
if HCN and CO are both incorporated into amorphous ice, and if the cometary material
was predominantly from regions where CO was more likely to be adsorbed onto forming
– 28 –
grains of amorphous water ice, then HCN depletion would indicate Kuiper Belt formation
distances, rather than closer-in Saturn distances.
5.2. Production Rate Ratios: CO2/CO
Another important marker in comae is the relative amount of CO2 and CO, which
is a useful test of thermal evolution and/or formation models (Feldman et al. 1997a;
Belton & Melosh 2009; A'Hearn et al. 2012). Although recent surveys show that CO2 >>
CO for most comets (Ootsubo et al. 2012; Reach et al. 2013b; Bauer et al. 2015b), CO
drives the activity for Hale-Bopp (beyond 4 AU) and 29P. Indeed, these two comets appear
to be "CO-rich" when compared to most other comets.
CO2 was directly detected in Hale-Bopp's coma at 4.3 microns using the Infrared
Space Observatory at 4.6 AU pre-perihelion and 4.9 AU post-perihelion (Crovisier 1997).
CO2 was also indirectly detected in Hale-Bopp with similar values via CO Cameron band
emission. These CO bands predominantly arise from CO2 dissociation (Weaver et al. 1997)
but also have contributions from electron impact excitation of CO (Bhardwaj & Raghuram
2011). CO2 production rates can be measured via high spectral resolution ground-based
observations of O[I] when cometary and telluric emission is significantly separated by
Doppler shift (McKay et al. 2012; Raghuram & Bhardwaj 2014; Decock et al. 2015), but
this technique was not used on Chiron, Hale-Bopp or 29P. Using directly measured CO
and CO2 production rates at approximately the same time from the literature, we calculate
production rate ratios of Q(CO2)/Q(CO) ∼ 0.1-0.3 for Hale-Bopp between 4-5 AU (Table
4). Searches for CO2 in 29P yielded upper limits of Q(CO2)/Q(CO) ∼ 0.01 (Ootsubo et al.
2012; Reach et al. 2013a), roughly ten times lower than for Hale-Bopp at ∼ 6 AU. Clearly,
at these large heliocentric distances, CO2 is not a significant driver for activity in either
Hale-Bopp or 29P. Interestingly, the Hale-Bopp and 29P Q(CO2)/Q(CO) ratios are
– 29 –
significantly lower than what is reported for many other periodic or dynamically new comets
Feldman et al. (1997b); A'Hearn et al. (2012); Bauer et al. (2015b); Reach et al. (2013b);
Ootsubo et al. (2012). It is important to note that most of these measurements of other
comets were obtained at heliocentric distances of 1 to 2.5 AU, which is much closer to the
Sun than the CO2 measurements in 29P and Hale-Bopp. It is possible that the production
rate ratios of CO2 and CO vary in distance from the Sun for an individual comet. The
production rate ratios may not reflect the nucleus abundances, but rather the activity
mechanism dominant at the time of observations. For example, at 29P's distance from the
Sun, the effective surface temperature (black-body) is roughly 120 K, which straddles the
threshold for water ice amorphous-crystalline transition. Due to differences in volatility
(sublimation pressures and rates), released CO will immediately flow outside, while released
CO2 would diffuse in the nucleus and may even re-condense at various depths. The presence
of ice and clathrate reservoirs, for CO2 and CO, could introduce alternative pathways.
However, the location of these components will have to be very close to the surface, within
∼1 skin depth, in order to match the quiescent activity of 29P.
Indeed, the measured Q(CO2)/Q(CO) ratio increased as Hale-Bopp approached the
Sun, which may be partly due to higher sublimation temperature for CO2 (see Table 1)
being reached by the comet. As part of a large infrared survey of comets that included 9
comets beyond 4 AU, Kelley et al. (2013) calculated each comet's efρ, which is a quantity
used to approximate the dust production rate. The study showed that log Q(CO2)/ǫfρ
broke into two groups of behavior for comets inside and outside 4 AU (see Figure 9 in
Bauer et al. (2015b)). The authors of that survey suggest that within 4 AU, CO2 activity
may be endogenic with the dust. This may be attributable to different source regions
(surface versus sub-surface) for cometary CO2 and CO emissions. Further searches for CO
and CO2 emission from other distantly active comets are needed to test models of volatile
production. Hale-Bopp and 29P appear to be CO-rich comets.
– 30 –
5.3. Spectral Line Profiles
Additional information is contained in spectral line profiles about how gas is emitted
from a comet nucleus. For example, consider that the millimeter wavelength spectral line
profiles of CO for Hale-Bopp (when beyond 4 AU), 29P and Chiron are all very narrow
(< 0.5 km/s), and sometimes are blue shifted by a small amount, ∼ 0.3-0.5 km s−1. The
numerical values of the velocity shifts are consistent with CO escaping from the nucleus at
velocities corresponding to hydrodynamical escape for v ∼ 0.4 km s−1 (Festou et al. 2001;
Gunnarsson et al. 2003). The spectra's blue-shifts indicate that the CO molecules' velocity
vector is primarily toward the Sun (and hence, the Earth, due to typically small phase
angles), and the emission can be understood if it occurs on the sunward side of the nucleus.
Such narrow lines indicate a narrow distribution of fluxes coming off the nucleus, which in
turn probably indicates a singular source for the CO, at least in terms of the temperature
and depth. There is a narrow range of thermal velocities that the CO molecules experience,
once released, and there is a specific length scale that these molecules traverse before leaving
the surface.
Sublimation from an icy grain halo is expected when an active comet is ∼ 4-6 AU from
the Sun. This occurred in Hale-Bopp's coma at ∼ 5 AU from the Sun, and there is strong
evidence that this develops in 29P and around other comets, such as Bowell, at 5 AU. This
is the likely origin of most of the CH3OH, OH, HCN, H2CO and H2S emission in Hale-Bopp
beyond 4 AU, as is illustrated in Figure 1. Interestingly, as Hale-Bopp moved closer to the
Sun (from 5 to 4 AU) the linewidths of methanol increased from ∼ 1 to 2 km s−1, while the
CO linewidths remained relatively constant, which further illustrates the differing behavior
for CO and CH3OH at large heliocentric distances (Womack et al. 1997).
– 31 –
5.4. A Consistent Set of CO Production Rates
Production rates for CO in these objects were published previously by several groups
over 20 years with different models and assumptions, including Senay & Jewitt (1994);
Crovisier et al. (1995); Womack et al. (1997); Womack & Stern (1999); Biver et al. (2002);
Gunnarsson et al. (2003); Paganini et al. (2013); DiSanti et al. (1999). Differences in
observing and modeling techniques make it difficult to assess similarities and differences of
the production rates among the comets. In order to better compare the CO production, and
probe the importance of nucleus size and thermal history, we derived the CO production
rates using our own data for Hale-Bopp, 29P and Chiron in a consistent manner. The data
for Hale-Bopp and Chiron were previously published using simpler models (Womack et al.
1997; Womack & Stern 1999). Most of the data for 29P are new and were obtained during
2016 Feb to May using the Arizona Radio Observatory 10-m Submillimeter Telescope
(SMT)(Wierzchos et al., in preparation). The 29P CO data are part of a separate project
and a full analysis will be published later. For now, we present the dates, heliocentric
distances and derived production rates in Table 5. One additional data point for 29P
was taken from Graham & Womack (1995). We assumed isotropic outgassing of CO from
the nucleus, rotational and excitation temperatures of Biver et al. (2002) and expansion
velocities from Gunnarsson et al. (2008). The excitation model includes both collisional
and fluorescence excitation in a manner similar to Crovisier & Le Bourlot 1983 and Biver
(1997). Applying a cone ejection model would reduce the numbers in Figure 2 by 40%, but
would not otherwise change the results. In Figure 2, we plot the new CO production rates
for Hale-Bopp, 29P and Chiron. The Hale-Bopp and 29P production rates are comparable
to values derived by Gunnarsson et al. (2003, 2008). In Figure 2 we also include Q(CO)
values for Hale-Bopp taken from Gunnarsson et al. (2003), to extend the heliocentric range
for Hale-Bopp; these values were not recalculated, because they used the same modeling
parameters as our new model. The upper limit for Chiron was taken from Drahus et al.
– 32 –
(2016).
The Hale-Bopp CO production rates in Figure 2 are all pre-perihelion and follow the
relationship Q(CO) = 3.5 × 1029r−2.2 molecules/second, where r is heliocentric distance
in astronomical units (AU). This heliocentric dependence agrees with what was found by
(Biver et al. 2002) for pre-perihelion data of Hale-Bopp. Later post-perihelion observations
of the comet out to ∼ 11 AU showed that Q(CO) followed a slightly less steep r−2
dependence (Biver et al. 2002; Gunnarsson et al. 2003). The pre-perihelion slope is steeper
than one would expect for a mechanism driven solely by insolation, which would produce a
r−2 dependence. The steeper slope may indicate that there is an additional source of energy
in this regime, such as crystallization of water ice. Based on the discussion and analysis
of Gunnarsson et al. (2003), we extrapolated the Q(CO) relationship for Hale-Bopp out to
11 AU, assuming an r−2 dependence to indicate the likely CO production rates at larger
heliocentric distances.
When adjusted for heliocentric distance, CO production rates in three distantly active
comets are remarkably uniform, despite different thermal processing histories (long-period
vs. Centaurs) and different nucleus sizes. We explore the apparently similar CO production
rates in the next section.
5.5. Observational Constraints to Existing Models of Distant Activity
The models that best explain the quiescent activity of Hale-Bopp and 29P are similar.
A porous comet nucleus comprising a mixture of ices and grains with a dust mantle at
the surface best fits the data obtained so far (e.g., Enzian et al. 1997; Enzian 1999;
Capria et al. 2002). In these models, crystallization of amorphous water ice near the surface
and subsequent release of trapped CO molecules is the dominant cause of activity for 29P
– 33 –
and Hale-Bopp within 7 AU. Activity in Hale-Bopp beyond 8 AU is more likely caused
by sublimation of near-surface CO ice, since the heating rate is too low for crystallization
to proceed in any appreciable rate. The survival of near-surface CO ice may be aided by
the low thermal conductivity maintained in very porous surface layers. Existing in-situ
measurements of comet nuclei point to a highly-porous bulk, as for the Rosetta and Deep
Impact measurements of comets 67P and 9P, respectively (Sierks et al. 2015, A'Hearn et
al. 2005). However, a porosity gradient has been inferred for 67P (from the CONSERT
sounding experiment), with an upper limit on near-surface porosity of 30In any case there
should exist a long-term reservoir for highly-volatile species, such as CO. The distinction
between the effects of an amorphous/crystalline water ice phase and a gradient in layering
and insulation falls outside the scope of this work and will be pursued by systematic
modeling efforts.
Moreover, the CO emission in 29P and Hale-Bopp has sunward velocity shifts at large
heliocentric distances. This implies that the CO released by both methods (sublimation of
CO ice and release of trapped CO during the crystallization of water ice) occurs from the
sub-solar regions of the nuclei. For Chiron, the surface temperatures are too warm for solid
CO or CO2 to be near the surface and sublimate, so instead, crystallization of water ice and
release of trapped gases is the more likely model (Prialnik et al. 1995; Capria et al. 2000).
A high obliquity for the rotation axis would allow Chiron's sub-solar regions to be heated
enough by the Sun to reach the measured color temperatures, which is also high enough for
crystallization to occur. These models also reproduce the quiescent dust production rates
of the comets.
However, we note that recent spacecraft data have revealed how geologically diverse
comets are. This surface geology and morphology evolution and the connection to the
sub-surface layers are a function of both origin and evolution. Thus, it is not just the specific
– 34 –
diffusivity and conductivity of the icy and silicate components, but also the configurations,
mixing ratios and dynamics of the different constituents that become important. Since
comets are significantly porous (e.g., Patzold et al. 2016; A'Hearn et al. 2011, 2005), the
way the porous medium evolves together with compositional and structural changes (phase
transitions, gas flow, cavities, small impacts, out-gassing) affects the determination of
bulk transport properties. Models of these interactions are inherently susceptible to some
degeneracy in the choice of parameters. The more chemical and physical information we
have about the nucleus and the behavior of the coma at various distances, the better we
can constrain the modeled profiles and compositions.
Gaseous outbursts can be explained if a region unusually rich in CO ice sublimates
when exposed to the Sun, or if a crystallization burst of water ice occurs. Either process
will cause an increase in CO and dust production (if there is much dust in that region).
Therefore, large dust jets emerge from cracks in the surface after being triggered by a
gaseous outburst, while the bulk of the observed CO (that generates the quiescent activity)
percolates through the comet surface that is at or near the sub-solar spot. Outbursts in
29P, in particular, are possibly triggered by surface erosion of the dust mantle (Enzian et al.
1997). For Chiron, sublimation of CO ice pockets is not likely, since, as discussed previously,
the surface temperatures are too warm for CO ice to be present. Instead, amorphous water
ice crystallization bursts are considered to be the most likely cause of outbursts for Chiron.
The mechanisms proposed to explain the distant activity in Hale-Bopp, 29P and
Chiron are possible in any comet which contains amorphous ice with trapped CO gas,
as well as frozen CO in their nuclei, as long as these materials are sufficiently near the
surface to be heated to the crystallization and/or sublimation temperatures. One might
wonder why most comets are not reported to be active until within 3 AU of the Sun. One
possibility is that most short-period comets exhausted their reservoirs of CO (and other
– 35 –
supervolatile ice) and converted their amorphous ice to the crystalline phase many millions
(or even billions) of years ago. Any remaining frozen CO and amorphous ice is too deeply
buried to be heated above the crystallization and sublimation temperatures during their
orbits. Another important point is that the nuclei of Hale-Bopp, 29P and Chiron are
larger than those of most typical comets, and this is probably responsible for much of the
distant activity. The larger volume exposes more material to heating, which provides more
opportunities for sublimation and crystallization. Some smaller, long-period comets may
exhibit CO production beyond 3 AU, but the quantities are too low to be detected with
current equipment.
Orbital history is thought to play a role in devolatilizing comet nuclei. Hale-Bopp's
orbit places it in the long-period comet category, which means it has experienced few heating
cycles over its lifetime, especially when compared to Chiron and 29P. It is dynamically
young and has probably experienced very different perihelion/aphelion evolution than
low-inclination bodies in the outer solar system (Bailey et al. 1996). As such, the reservoirs
of volatile species in the nucleus have been processed in a different manner and could
have remained more accessible to subsequent apparition. Large gas fluxes from previously
untapped reservoirs of volatile species, which can efficiently entrain and drag dust grains
from various depths in the nucleus, could have been the main source of the distinctive
brightness of Hale-Bopp. The highly-volatile nature of the gas (and perhaps larger
abundances of CO) may have been the main cause for very distant activity, both pre- and
post-perihelion. 29P orbits much closer to the Sun, at r ∼ 6 AU, and thermal evolution
from previous orbits probably depleted any highly-volatile solid ice from at least the first
100 or so meters. Accessible solid ice layers are probably rare, and thus not the main driver
of activity for 29P. This is also probably true, to some extent, for Chiron, which orbits
between 8.5 - 13 AU. By this measure alone, Hale-Bopp should have produced significantly
more CO than 29P or Chiron. However, as Figure 2 shows, the CO production rates are
– 36 –
consistent between the comets. 29P produces CO at approximately the same rate as was
observed in Hale-Bopp when it was at r ∼ 6 AU. Hale-Bopp's and 29P's very different
orbital histories appear not to be reflected in their CO production rates at 6 AU. Chiron
is a little more complicated to sort out. The CO emission detection in Chiron at 8.5 AU
is consistent with the CO production rate in Hale-Bopp at about the same heliocentric
distance post-perihelion. However, it is important to keep in mind that other searches for
CO in Chiron led to upper limits which were significantly lower than the detection, which
presumably occurred during an outburst. The non-detection limit made over two years
implies that if Chiron does continually emit CO, it is likely at a much lower rate than
Hale-Bopp or 29P. Dust production rates estimated from optical-infrared images are also
consistent with significantly lower activity overall compared to Hale-Bopp at this distance
(Romon-Martin et al. 2003; Weiler et al. 2003). Given the low dust production rates,
Chiron is not likely to be outgassing significant amounts of another volatile, such as CO2.
The Chiron data are consistent with very little constant volatile outgassing punctuated by
sporadic outbursts.
Although Hale-Bopp's long orbital period means that it is substantially less
heated/processed than short-period comets, it has endured several passages near the Sun,
which may be evident in comparisons with a true dynamically new comet, making its
first perihelion approach. Detailed comparisons of observations of Hale-Bopp with such
dynamically new comets at large distances would be very useful and should be encouraged.
Currently, very few direct comparisons are possible beyond ∼ 2 AU, but one recently
studied dynamically new comet, C/2009 P1 Garradd, showed a striking difference. Garradd
did not show the strong correlation between CO activity and heliocentric distance as
Hale-Bopp. Rather than show a symmetric increase and decrease of CO around perihelion,
Garradd's CO production continued to increase throughout the entire time period that
could be studied, out to ∼ 2-3 AU post-perihelion (Feaga et al. 2014; McKay et al. 2015).
– 37 –
Unfortunately, observations of CO emission beyond 4 AU do not exist for comet Garradd.
It would be particularly valuable to compare CO behavior for these two classes of comets
when they are beyond the water-ice sublimation limit. Also, closer to the Sun, CO may be
produced in significant amounts as daughter products from parent species. Such competing
sources for CO production near the Sun complicates efforts to account for how much CO is
natal and driving the long-term activity.
Of the three objects discussed in this paper, two that have very different orbital
histories (29P and Hale-Bopp) and thus should have very different CO production rates, in
fact have very similar CO production rates. In addition, two which are classified as Centaurs
with relatively close-in orbits (29P and Chiron) are expected to produce similar amounts
of CO had they started with similar compositions. However, they have vastly different CO
production rates. This is a very small sample set, obviously, but from studying these three
objects we conclude that orbital history alone cannot explain the CO production rates in
these objects. Further constraining measurements of CO emission from other Centaurs are
needed to explore the role CO outgassing plays in Centaurs.
5.6. CO Output Adjusted For Nucleus Size
A comet's size is also expected to be correlated with how productive it is when it is
heated by the Sun. For larger objects, once amorphous ice starts transitioning, triggered by
small increases in temperature (from insolation, impacts, chemical reactions, etc.), it will
depend only on the heat wave coming in from the surface. This heat wave is just a function
of heliocentric distance, if we assume all other properties are similar (material, porosity,
distribution, etc.).
Larger comets are more susceptible to radiogenic heating, which may lead to
– 38 –
devolatilization of the original ice. As discussed earlier, comets smaller than 10 km probably
have their initial temperature preserved deeper than about 100 meters. However, 29P,
Chiron and Hale-Bopp all have radii larger than 10 km, and so probably experienced some
amount of radiogenic heating. Chiron is much larger than 29P and Hale-Bopp and so
probably underwent even more radiogenic heating (unless 29P and Hale-Bopp are fragments
of larger objects). Consequently, on the basis of size, one might expect Chiron to be more
processed, and thus have less outgassing than 29P or Hale-Bopp. On the other hand, if one
assumes comet nuclei to be roughly heterogeneous, then bigger comets should have higher
outputs when heated.
CO emission was searched for, and not found, in a collection of KBO's and Centaurs
by Bockel´ee-Morvan et al. (2001); Jewitt et al. (2008); Drahus et al. (2016). Many of the
Centaurs were first observed by Bockel´ee-Morvan et al. (2001) and Jewitt et al. (2008) and
then re-analyzed by Drahus et al. (2016), who also made new observations. The authors of
the first survey proposed that if the objects' CO activity followed the trend of Hale-Bopp's
gas activity with heliocentric distance, and is proportional to the object's diameter, D, then
the relationship Q(CO) ∼ r−2 × D2 should hold, and CO should have been detected in these
very distant objects. The upper limits they obtained are significantly below what would
be predicted from this simple scaling law, which they conclude is evidence that that these
objects underwent significant CO devolatilization since their formation (De Sanctis et al.
2001b). The authors of the second survey drew the same conclusions about Centaurs likely
being depleted in CO using similar arguments, whereas the third survey raised the question
of whether CO outgassing should be correlated with heliocentric distance or size, but
instead a small heliocentric distance might be more important. 29P is a notable exception
to the list of apparently CO-depleted Centaurs, and its nearly circular orbit has a relatively
small heliocentric distance of ∼ 6 AU.
– 39 –
To explore this further, we calculated CO "specific production rates," which are the
production rates divided by the square of the nucleus diameter. Specific production rates
were calculated for Hale-Bopp, 29P and Chiron using CO production rates from Figure 2.
Note that the Chiron upper limit at ∼ 9.5 AU is from Drahus et al. (2016), which they
recalculated using Bockel´ee-Morvan et al. (2001) data. The specific production rates are
plotted in Figure 4. Accordingly, Chiron is approximately 5-15 times depleted in CO when
compared to Hale-Bopp, a proxy for a relatively unprocessed comet.4 We do not have our
own Hale-Bopp data at 8-10 AU, and hence we extrapolated Q(CO) from the r−2 fit to the
Q(CO) data, which is also consistent with Q(CO) values from Gunnarsson et al. (2003).
We conclude that there is measurably less CO produced per surface area from Chiron than
either Hale-Bopp or 29P. This is consistent with radiogenic heating devolatilization being
an important parameter in nucleus modeling, since Chiron is much larger than either of the
other two comets.
As discussed in the previous section, Hale-Bopp and 29P have very similar CO
production rates, even when corrected for nucleus size. We see no evidence for
devolatilization for 29P when compared to Hale-Bopp; in fact, when outbursting, 29P
outproduces Hale-Bopp by almost a factor of ten. This may point to 29P being a more
recent entrant to the inner solar system than Chiron.
Since Chiron does not get closer than ∼ 8.5 AU, where Tbb ∼ 100K, it can maintain a
4A similar calculation was performed by Bockel´ee-Morvan et al. (2001) and found Chiron
to be depleted by a factor of ∼ 35 with respect to Hale-Bopp. This difference in depletion
values can be largely explained by different values used for Hale-Bopp's nucleus diameter,
D, and CO production rate at 8-9 AU. For example, Bockel´ee-Morvan et al. (2001) assumed
D=40 km diameter (we use 60 km) and they used an older CO production rate value for
Hale-Bopp that was later updated by Gunnarsson et al. (2003), which we also use.
– 40 –
slow crystallization rate of ∼ 1000 years (and rate of trapped volatile release), if we assume
maximum efficiency. Perhaps Chiron never experienced large temperature excursions
to its surface, since the perihelion was never much inside of its current one. However,
by considering the noted higher color temperature and low CO production rate, we
find it consistent with Chiron's being appreciably devolatilized over a fresher object like
Hale-Bopp. Also, Chiron's interior may be warmer due to early radiogenic heating and slow
cooling rates, as it is a larger object than both Hale-Bopp and 29P.
6. Conclusions
All Hale-Bopp, 29P and Chiron have in common that: 1) They exhibit long-term
quiescent activity beyond 4 AU that is punctuated by sporadic outbursts; 2) Emission from
CN and CO was seen in their coma; 3) The CO line profiles were narrow, ∆v(FWHM) ∼
0.3-0.5 km s−1; 4) They have spent most of their lifetime far from the Sun;5 5) They have
larger-than-usual nuclei.
There are notable differences among the comets, including: 1) The CO emission
had blue-shifted velocity components of ∼ -0.4 km/s (sunward) for 29P and Hale-Bopp,
but almost no shift for Chiron's emission; 2) CO was seen only once in Chiron, but was
repeatedly observable for long-term study in Hale-Bopp and 29P (and 29P is observable,
even now); 3) Hale-Bopp had much higher dust production rates than seen in 29P or
Chiron.
Sublimation or escape of molecules from a grain halo at ∼ 5-6 AU appears to be a
5Chiron and 29P have only recently entered their current orbits and may be precursors
to Jupiter Family comets (Fernandez & Gallardo 1994), and Hale-Bopp has an aphelion of
∼ 360 AU, spends most of its time at very large heliocentric distances.
– 41 –
typical development in distantly active comets, and may be the main source of emission
of OH, HCN, CH3OH, H2CO, and H2S in Hale-Bopp between 3 – 6 AU. This may also
explain the origin of OH and CN molecules in other distant comets. Spectral line profiles
and spatial mapping should be used whenever possible to determine a nuclear or extended
origin for volatiles in distant comets. Observed radio line profiles are consistent with
the development and sublimation of icy grains in the coma at about 5-6 AU, and this is
probably a common feature in distantly active comets, and an important source of other
volatiles within 5 AU. Note that narrow CO line profiles for all three main bodies in this
study indicate a nuclear origin for CO beyond ∼ 4 AU, and not an extended source, as with
these other molecular cases.
Our calculated CN/CO and CO2/CO production rate ratios for Hale-Bopp and 29P
show that HCN and CO2 were both produced in significantly smaller amounts than CO,
and can only be minor contributors to distant activity of these two objects. For Chiron, no
limits exist for CO2 emission, and CN and CO were both detected only once at the ∼ 3.5σ
level. The CN measurements are consistent with HCN being an insignificant contributor
to Chiron's coma. The CO data correspond to a gas production rate high enough to drive
the observed dust coma activity; however, the CO detected may have also resulted from
an outburst and we cannot rule out significant contributions from undetected CO2 or
other supervolatile gases. We point out that the CN/CO production rate ratios that we
derive from the literature are remarkably similar in all three distantly active comets, which
warrants follow-up study.
When adjusted for heliocentric distance, our analysis shows that CO production
rates in three distantly active comets are remarkably uniform, despite different thermal
processing histories (long-period vs. Centaurs) and different nucleus sizes. Specific CO
production rates, Q(CO)/D2, were calculated in order to take into account nucleus size,
– 42 –
which is predicted to play a large role in cometary activity. We find that orbital history
does not appear to play a significant role in explaining the CO production rates for 29P
and Hale-Bopp, which confirms the conclusion in A'Hearn et al. (2012) for comets within
2 AU. 29P outproduces Hale-Bopp at the same heliocentric distance, although it has been
subjected to solar heating far longer, and models predict it should have been devolatilized
over a fresher comet, like Hale-Bopp. However, we may see evidence for the relevance
of nucleus size when considering specific production rates. If CO is present in Chiron's
coma on a long-term basis, it must be in relatively small amounts and consistent with a
CO depletion by at least 5-15 times over Hale-Bopp, perhaps due to increased radiogenic
heating made possible by its much larger size or its higher processing due to orbital history.
The model which best reproduces the distant activity of 29P and Hale-Bopp is a comet
nucleus composed of dust, amorphous water ice and smaller amounts of trapped CO and
CO2 trapped in the amorphous ice and a build-up of a thin porous dust mantle (Prialnik
1997a,b). Within ∼ 4 - 7 AU, the water ice undergoes crystallization near the surface and
releases trapped gases and dust. There is a large range of free parameters in the models.
The spin state and specific porous medium properties play a large role in the differences in
activity between comets. The release of CO and subsequent permeability of the gas flow
could inhibit emission, especially if the local temperature (and pressure) gradients vary.
This variation could be a strong function of solar radiation heating and surface cooling,
which are functions of the spin state and surface morphology. Since Rosetta, the latter has
become an increasingly important factor (Thomas et al. 2015).
Models for Chiron's distant activity are presented in Prialnik et al. (1995); Capria et al.
(2000); Fanale & Salvail (1997). The first two include amorphous ice and rely on release of
trapped CO, while the latter starts with a layered model of dust mantle, thick water ice layer
and CO layer underneath. The amorphous ice models better reproduce the distant activity,
– 43 –
although they are constrained by a lack of observations of CO and dust production. The
latter is a convolution of gas fluxes, dust properties and coma behavior. Models dominated
by crystalline water ice and CO/CO2 ices are not successful (Prialnik et al. 1995). Instead,
Chiron's activity appears to be best explained by a release of trapped volatile species
in the water ice matrix. However, a more complete understanding of the outburst or
quiescent nature of this activity would require more measurements of CO production
rates, at different times. An optimal addition would be observations of other species (e.g.,
methanol, methane, CO2 , etc.), which would enable more stringent constraints on the
nucleus' composition ratios and thermal history. Radiogenic heating may be important for
explaining Chiron's low-level activity that appears CO-depleted. Furthermore, if rings are
confirmed around Chiron, then previously reported long-term changes in visual magnitude
and infrared spectral ice features should be re-examined.
7. Acknowledgements
The authors thank the anonymous referee for useful comments and feedback, which
improved the paper. This material is based upon work supported by (while MW was
serving at) the National Science Foundation (NSF). MW also acknowledges support from
NSF grants AST-1009933 and AST-1615917. Any opinion, findings, and conclusions or
recommendations expressed in this material are those of the author(s) and do not necessarily
reflect the views of the National Science Foundation. The SMT is operated by the Arizona
Radio Observatory (ARO), the Steward Observatory, and the University of Arizona, with
support through the NSF University Radio Observatories program (AST-1140030).
– 44 –
REFERENCES
A'Hearn, M. F., Belton, M. J. S., Delamere, W. A., Feaga, L. M., Hampton, D., Kissel,
J., Klaasen, K. P., McFadden, L. A., Meech, K. J., Melosh, H. J., Schultz, P. H.,
Sunshine, J. M., Thomas, P. C., Veverka, J., Wellnitz, D. D., Yeomans, D. K., Besse,
S., Bodewits, D., Bowling, T. J., Carcich, B. T., Collins, S. M., Farnham, T. L.,
Groussin, O., Hermalyn, B., Kelley, M. S., Kelley, M. S., Li, J.-Y., Lindler, D. J.,
Lisse, C. M., McLaughlin, S. A., Merlin, F., Protopapa, S., Richardson, J. E., &
Williams, J. L. 2011, Science, 332, 1396
A'Hearn, M. F., Belton, M. J. S., Delamere, W. A., Kissel, J., Klaasen, K. P., McFadden,
L. A., Meech, K. J., Melosh, H. J., Schultz, P. H., Sunshine, J. M., Thomas, P. C.,
Veverka, J., Yeomans, D. K., Baca, M. W., Busko, I., Crockett, C. J., Collins, S. M.,
Desnoyer, M., Eberhardy, C. A., Ernst, C. M., Farnham, T. L., Feaga, L., Groussin,
O., Hampton, D., Ipatov, S. I., Li, J.-Y., Lindler, D., Lisse, C. M., Mastrodemos, N.,
Owen, W. M., Richardson, J. E., Wellnitz, D. D., & White, R. L. 2005, Science, 310,
258
A'Hearn, M. F., Feaga, L. M., Keller, H. U., Kawakita, H., Hampton, D. L., Kissel, J.,
Klaasen, K. P., McFadden, L. A., Meech, K. J., Schultz, P. H., Sunshine, J. M.,
Thomas, P. C., Veverka, J., Yeomans, D. K., Besse, S., Bodewits, D., Farnham,
T. L., Groussin, O., Kelley, M. S., Lisse, C. M., Merlin, F., Protopapa, S., &
Wellnitz, D. D. 2012, ApJ, 758, 29
A'Hearn, M. F., Millis, R. C., Schleicher, D. O., Osip, D. J., & Birch, P. V. 1995, Icarus,
118, 223
A'Hearn, M. F., Schleicher, D. G., Millis, R. L., Feldman, P. D., & Thompson, D. T. 1984,
AJ, 89, 579
– 45 –
Bar-Nun, A., Kleinfeld, I., & Kochavi, E. 1988, Phys. Rev. B, 38, 7749
Bar-Nun, A. & Owen, T. 1998, in Astrophysics and Space Science Library, Vol. 227, Solar
System Ices, ed. B. Schmitt, C. de Bergh, & M. Festou, 353
Bauer, J. M., Stevenson, R., Kramer, E., Mainzer, A. K., Grav, T., Masiero, J. R.,
Fern´andez, Y. R., Cutri, R. M., Dailey, J. W., Masci, F. J., Meech, K. J., Walker,
R., Lisse, C. M., Weissman, P. R., Nugent, C. R., Sonnett, S., Blair, N., Lucas, A.,
McMillan, R. S., Wright, E. L., WISE, t., & NEOWISE Teams. 2015a, ApJ, 814, 85
-. 2015b, ApJ, 814, 85
Belton, M. J. S. & Melosh, J. 2009, Icarus, 200, 280
Bhardwaj, A. & Raghuram, S. 2011, MNRAS, 412, L25
Biver. 1997, Universit Paris 7
Biver, N., Bockelee-Morvan, D., Colom, P., Crovisier, J., Davies, J. K., Dent, W. R. F.,
Despois, D., Gerard, E., Lellouch, E., Rauer, H., Moreno, R., & Paubert, G. 1997a,
Science, 275, 1915
Biver, N., Bockel´ee-Morvan, D., Colom, P., Crovisier, J., Germain, B., Lellouch, E., Davies,
J. K., Dent, W. R. F., Moreno, R., Paubert, G., Wink, J., Despois, D., Lis, D. C.,
Mehringer, D., Benford, D., Gardner, M., Phillips, T. G., Gunnarsson, M., Rickman,
H., Winnberg, A., Bergman, P., Johansson, L. E. B., & Rauer, H. 1997b, Earth
Moon and Planets, 78, 5
Biver, N., Bockel´ee-Morvan, D., Colom, P., Crovisier, J., Henry, F., Lellouch, E., Winnberg,
A., Johansson, L. E. B., Gunnarsson, M., Rickman, H., Rantakyro, F., Davies, J. K.,
Dent, W. R. F., Paubert, G., Moreno, R., Wink, J., Despois, D., Benford, D. J.,
– 46 –
Gardner, M., Lis, D. C., Mehringer, D., Phillips, T. G., & Rauer, H. 2002, Earth
Moon and Planets, 90, 5
Biver, N., Bockel´ee-Morvan, D., Crovisier, J., Davies, J. K., Matthews, H. E., Wink, J. E.,
Rauer, H., Colom, P., Dent, W. R. F., Despois, D., Moreno, R., Paubert, G., Jewitt,
D., & Senay, M. 1999, AJ, 118, 1850
Biver, N., Rauer, H., Despois, D., Moreno, R., Paubert, G., Bockel´ee-Morvan, D., Colom,
P., Crovisier, J., G´erard, E., & Jorda, L. 1996, Nature, 380, 137
Bockelee-Morvan, D., Biver, N., Crovisier, J., de Val-Borro, M., Fulton, T., Hartogh, P.,
Hutsem´ekers, D., Jarchow, C., Jehin, E., Kidger, M., Kueppers, M., Lellouch, E.,
Lis, D., Manfroid, J., Moreno, R., Rengel, M., Swinyard, B. C., Szutowicz, S.,
Vandenbussche, B., & HssO Team. 2010, in Bulletin of the American Astronomical
Society, Vol. 42, AAS/Division for Planetary Sciences Meeting Abstracts #42, 946
Bockel´ee-Morvan, D., Biver, N., Jehin, E., Cochran, A. L., Wiesemeyer, H., Manfroid, J.,
Hutsem´ekers, D., Arpigny, C., Boissier, J., Cochran, W., Colom, P., Crovisier, J.,
Milutinovic, N., Moreno, R., Prochaska, J. X., Ramirez, I., Schulz, R., & Zucconi,
J.-M. 2008, ApJ, 679, L49
Bockelee-Morvan, D., Biver, N., Opitom, C., Hutsemekers, D., Crovisier, J., Jehin, E.,
Hartogh, P., Szutowizc, S., Lellouch, E., Kidger, M., Vandenbussche, B., Zakharov,
V., & HSSO Team. 2014, in Asteroids, Comets, Meteors 2014, ed. K. Muinonen,
A. Penttila, M. Granvik, A. Virkki, G. Fedorets, O. Wilkman, & T. Kohout
Bockel´ee-Morvan, D., Lellouch, E., Biver, N., Paubert, G., Bauer, J., Colom, P., & Lis,
D. C. 2001, Astron. and Astrophy., 377, 343
Boehnhardt, H., Birkle, K., Fiedler, A., Jorda, L., Thomas, N., Peschke, S., Rauer, H.,
– 47 –
Schulz, R., Schwehm, G., Tozzi, G., & West, R. 1997, Earth Moon and Planets, 78,
179
Boice, D. C. & Huebner, W. F. 1993, in Bulletin of the American Astronomical Society,
Vol. 25, AAS/Division for Planetary Sciences Meeting Abstracts #25, 1064
Boss, A. P. 1994, Icarus, 107, 422
Braunstein, M., Comstock, R., Hoffman, P., Womack, M., Deglman, F., Pinnick, D., Aaker,
G., Goldschen, M., Jacobson, A., Zilka, J., Faith, D., Moore, S., Ricotta, J., Weist,
A., & Modi, C. 1997, Earth Moon and Planets, 78, 219
Brown, G. N. & Ziegler, W. T. 1979, Adv. Cryogen. Engineering, 25, 662
Buratti, B. J. & Dunbar, R. S. 1991, ApJ, 366, L47
Bus, S. J., A'Hearn, M. F., Schleicher, D. G., & Bowell, E. 1991, Science, 251, 774
Bus, S. J., Bowell, E., Harris, A. W., & Hewitt, A. V. 1989, Icarus, 77, 223
Bus, S. J., Buie, M. W., Schleicher, D. G., Hubbard, W. B., Marcialis, R. L., Hill, R.,
Wasserman, L. H., Spencer, J. R., Millis, R. L., Franz, O. G., Bosh, A. S., Dunham,
E. W., Ford, C. H., Young, J. W., Elliott, J. L., Meserole, R., Olkin, C. B.,
McDonald, S. W., Foust, J. A., Sopata, L. M., & Bandyopadhyay, R. M. 1996,
Icarus, 123, 478
Bus, S. J., Stern, S. A., & A'Hearn, M. F. 1993, Proceedings of the Distant Comet Activity
Workshop, 41
Campins, H., Telesco, C. M., Osip, D. J., Rieke, G. H., Rieke, M. J., & Schulz, B. 1994, AJ,
108, 2318
Capria, M. T., Coradini, A., & de Sanctis, M. C. 2002, Earth Moon and Planets, 90, 217
– 48 –
Capria, M. T., Coradini, A., De Sanctis, M. C., & Orosei, R. 2000, AJ, 119, 3112
Cochran, A. L. 2002, ApJ, 576, L165
Cochran, A. L., Barker, E. S., & Cochran, W. D. 1980, Astron. J., 85, 474
Cochran, A. L. & Cochran, W. D. 1991, Icarus, 90, 172
Cochran, A. L., Cochran, W. D., & Barker, E. S. 1982, ApJ, 254, 816
-. 2000, Icarus, 146, 583
Cook, J. C., Desch, S. J., & Wyckoff, S. 2005, in Bulletin of the American Astronomical
Society, Vol. 37, AAS/Division for Planetary Sciences Meeting Abstracts #37, 645
Crifo, J. F., Rodionov, A. V., & Bockel´ee-Morvan, D. 1999, Icarus, 138, 85
Crovisier, J. 1997, Earth Moon and Planets, 79, 125
Crovisier, J., Biver, N., Bockelee-Morvan, D., Colom, P., Jorda, L., Lellouch, E., Paubert,
G., & Despois, D. 1995, Icarus, 115, 213
Crovisier, J. & Le Bourlot, J. 1983, A&A, 123, 61
Crovisier, J., Leech, K., Bockel´ee-Morvan, D., Brooke, T. Y., Hanner, M. S., Altieri, B.,
Keller, H. U., & Lellouch, E. 1997, in ESA Special Publication, Vol. 419, The first
ISO workshop on Analytical Spectroscopy, ed. A. M. Heras, K. Leech, N. R. Trams,
& M. Perry, 137
De Sanctis, M. C., Capria, M. T., & Coradini, A. 2001a, AJ, 121, 2792
-. 2001b, AJ, 121, 2792
Decock, A., Jehin, E., Rousselot, P., Hutsem´ekers, D., Manfroid, J., Raghuram, S.,
Bhardwaj, A., & Hubert, B. 2015, A&A, 573, A1
– 49 –
Dello Russo, N., Kawakita, H., Vervack, R. J., & Weaver, H. A. 2016, Icarus, 278, 301
Delsemme, A. H. & Miller, D. C. 1971, Planet. Space Sci., 19, 1229
Despois, D., Biver, N., Bockel´ee-Morvan, D., & Crovisier, J. 2005, in IAU Symposium, Vol.
231, Astrochemistry: Recent Successes and Current Challenges, ed. D. C. Lis, G. A.
Blake, & E. Herbst, 469–478
DiSanti, M., Mumma, N. J., Dello Russo, N., Magee-Sauer, K., Novak, R., & Rettig, T. W.
1999, Nature, 399, 662
Dones, L., Brasser, R., Kaib, N., & Rickman, H. 2015, Space Sci. Rev., 197, 191
Drahus, M., Yang, B., Lis, D. C., & Jewitt, D. 2016, MNRAS
Duffard, R., Lazzaro, D., Pinto, S., Carvano, J., Angeli, C., Alvarez-Candal, A., &
Fern´andez, S. 2002, Icarus, 160, 44
Duffard, R., Pinilla-Alonso, N., Ortiz, J. L., Alvarez-Candal, A., Sicardy, B., Santos-Sanz,
P., Morales, N., Colazo, C., Fern´andez-Valenzuela, E., & Braga-Ribas, F. 2014,
A&A, 568, A79
Durda, D. D. & Stern, S. A. 2000, Icarus, 145, 220
Elliot, J. L., Olkin, C. B., Dunham, E. W., Ford, C. H., Gilmore, D. K., Kurtz, D., Lazzaro,
D., Rank, D. M., Temi, P., Bandyopadhyay, R. M., Barroso, J., Barucci, A., Bosh,
A. S., Buie, M. W., Bus, S. J., Dahn, C. C., Foryta, D. W., Hubbard, W. B., Lopes,
D. F., Marcialis, R. L., McDonald, S. W., Millis, R. L., Reitsema, H., Schleicher,
D. G., Sicardy, B., Stone, R. P. S., & Wasserman, L. H. 1995, Nature, 373, 46
Enzian, A. 1999, Space Sci. Rev., 90, 131
Enzian, A., Cabot, H., & Klinger, J. 1997, A&A, 319, 995
– 50 –
Fanale, F. P. & Salvail, J. R. 1997, Icarus, 125, 397
Farinella, P. & Davis, D. R. 1996, Science, 273, 938
Feaga, L. M., A'Hearn, M. F., Farnham, T. L., Bodewits, D., Sunshine, J. M., Gersch,
A. M., Protopapa, S., Yang, B., Drahus, M., & Schleicher, D. G. 2014, AJ, 147, 24
Feldman, P. D., Cochran, A. L., & Combi, M. R. Spectroscopic investigations of fragment
species in the coma, ed. M. C. Festou, H. U. Keller, & H. A. Weaver, 425–447
Feldman, P. D., Festou, M. C., Tozzi, P., & Weaver, H. A. 1997a, ApJ, 475, 829
-. 1997b, ApJ, 475, 829
Feldman, P. D., McPhate, J. B., Weaver, H. A., Tozzi, G.-P., & A'Hearn, M. F. 1996, in
Bulletin of the American Astronomical Society, Vol. 28, AAS/Division for Planetary
Sciences Meeting Abstracts #28, 1084
Fernandez, J. A. 1980, Icarus, 42, 406
Fernandez, J. A. & Gallardo, T. 1994, A&A, 281, 911
Fern´andez, Y. R. 2000, Earth Moon and Planets, 89, 3
-. 2009, Planet. Space Sci., 57, 1218
Fern´andez, Y. R., Wellnitz, D. D., Buie, M. W., Dunham, E. W., Millis, R. L., Nye, R. A.,
Stansberry, J. A., Wasserman, L. H., A'Hearn, M. F., Lisse, C. M., Golden, M. E.,
Person, M. J., Howell, R. R., Marcialis, R. L., & Spitale, J. N. 1999, Icarus, 140, 205
Festou, M. C., Barale, O., Davidge, T., Stern, S. A., Tozzi, G. P., Womack, M., & Zucconi,
J. M. 1998, in Bulletin of the American Astronomical Society, Vol. 30, AAS/Division
for Planetary Sciences Meeting Abstracts #30, 1089
– 51 –
Festou, M. C., Gunnarsson, M., Rickman, H., Winnberg, A., & Tancredi, G. 2001, Icarus,
150, 140
Fitzsimmons, A. & Cartwright, I. M. 1996, MNRAS, 278, L37
Fornasier, S., Lellouch, E., Muller, T., Santos-Sanz, P., Panuzzo, P., Kiss, C., Lim, T.,
Mommert, M., Bockel´ee-Morvan, D., Vilenius, E., Stansberry, J., Tozzi, G. P.,
Mottola, S., Delsanti, A., Crovisier, J., Duffard, R., Henry, F., Lacerda, P., Barucci,
A., & Gicquel, A. 2013, A&A, 555, A15
Foster, M. J., Green, S. F., McBride, N., & Davies, J. K. 1999, Icarus, 141, 408
Fox, J. L. & Black, J. H. 1989, Geophys. Res. Lett., 16, 291
Fray, N., B´enilan, Y., Cottin, H., Gazeau, M.-C., & Crovisier, J. 2005, Planet. Space Sci.,
53, 1243
Fulle, M. 1992, Nature, 359, 42
-. 1994, A&A, 282, 980
Ghormley, J. A. 1968, J. Chem. Phys., 48, 503
Graham, Jr., R. A. & Womack, M. 1995, in B.A.A.S., Vol. 27, #186, 857
Gronkowski, P. & Smela, J. 1998, A&A, 338, 761
Groussin, O., Lamy, P., & Jorda, L. 2004, A&A, 413, 1163
Gunnarsson, M., Bockel´ee-Morvan, D., Biver, N., Crovisier, J., & Rickman, H. 2008,
Astron. and Astroph., 484, 537
Gunnarsson, M., Bockel´ee-Morvan, D., Winnberg, A., Rickman, H., Crovisier, J., Biver, N.,
Colom, P., Davies, J. K., Despois, D., Henry, F., Johansson, L. E. B., Moreno, R.,
Paubert, G., & Rantakyro, F. T. 2003, A&A, 402, 383
– 52 –
Gunnarsson, M., Rickman, H., Festou, M. C., Winnberg, A., & Tancredi, G. 2002, Icarus,
157, 309
Hahn, G. & Bailey, M. E. 1990, Nature, 348, 132
Hanner, M. S. 1981, Icarus, 47, 342
-. 1984, ApJ, 277, L75
Hartmann, W. K., Tholen, D. J., Meech, K. J., & Cruikshank, D. P. 1990, Icarus, 83, 1
Haruyama, J., Yamamoto, T., Mizutani, H., & Greenberg, J. M. 1993, J. Geophys. Res.,
98, 15
Hassig, M., Altwegg, K., Balsiger, H., Bar-Nun, A., Berthelier, J. J., Bieler, A., Bochsler,
P., Briois, C., Calmonte, U., Combi, M., De Keyser, J., Eberhardt, P., Fiethe, B.,
Fuselier, S. A., Galand, M., Gasc, S., Gombosi, T. I., Hansen, K. C., Jackel, A.,
Keller, H. U., Kopp, E., Korth, A., Kuhrt, E., Le Roy, L., Mall, U., Marty, B.,
Mousis, O., Neefs, E., Owen, T., R`eme, H., Rubin, M., S´emon, T., Tornow, C.,
Tzou, C.-Y., Waite, J. H., & Wurz, P. 2015, Science, 347, aaa0276
Hughes, D. W. 1991, in Astrophysics and Space Science Library, Vol. 167, IAU Colloq. 116:
Comets in the post-Halley era, ed. R. L. Newburn, Jr., M. Neugebauer, & J. Rahe,
825–851
Iro, N., Gautier, D., Hersant, F., Bockel´ee-Morvan, D., & Lunine, J. I. 2003, Icarus, 161,
511
Irvine, W. M., Schloerb, F. P., Crovisier, J., Fegley, Jr., B., & Mumma, M. J. 2000,
Protostars and Planets IV, 1159
Ivanova, A., Afanasiev, V., Korsun, P., Baransky, A., Andreev, M., & Ponomarenko, V.
2015, Highlights of Astronomy, 16, 176
– 53 –
Ivanova, O. V., Luk'yanyk, I. V., Kiselev, N. N., Afanasiev, V. L., Picazzio, E., Cavichia,
O., de Almeida, A. A., & Andrievsky, S. M. 2016, Planet. Space Sci., 121, 10
Jewitt, D. 1990, Astroph. J., 351, 277
-. 1993, Proceedings of the Workshop on the Activity of Distant Comets, 64
Jewitt, D., Garland, C. A., & Aussel, H. 2008, AJ, 135, 400
Jewitt, D., Senay, M., & Matthews, H. 1996, Science, 271, 1110
Kelley, M. S., Lindler, D. J., Bodewits, D., A'Hearn, M. F., Lisse, C. M., Kolokolova, L.,
Kissel, J., & Hermalyn, B. 2013, Icarus, 222, 634
Klinger, J. 1980, Science, 209, 634
-. 1981, Icarus, 47, 320
Kouchi, A., Greenberg, J. M., Yamamoto, T., & Mukai, T. 1992, ApJ, 388, L73
Kouchi, A. & Sirono, S.-i. 2001, Geophys. Res. Lett., 28, 827
Kouchi, A., Yamamoto, T., Kozasa, T., Kuroda, T., & Greenberg, J. M. 1994, A&A, 290
Kulyk, I., Korsun, P., Rousselot, P., Afanasiev, V., & Ivanova, O. 2016, Icarus, 271, 314
Lamy, P. L. 1974, A&A, 35, 197
Larson, S. M. 1980, ApJ, 238, L47
Laufer, D., Kochavi, E., & Bar-Nun, A. 1987, Phys. Rev. B, 36, 9219
Lazzaro, D., Florczak, M. A., Angeli, C. A., Carvano, J. M., Betzler, A. S., Casati, A. A.,
Barucci, M. A., Doressoundiram, A., & Lazzarin, M. 1997, Planet. Space Sci., 45,
1607
– 54 –
Lebofsky, L. A., Tholen, D. J., Rieke, G. H., & Lebofsky, M. J. 1984, Icarus, 60, 532
Lellouch, E., Crovisier, J., Lim, T., Bockelee-Morvan, D., Leech, K., Hanner, M. S., Altieri,
B., Schmitt, B., Trotta, F., & Keller, H. U. 1998, A&A, 339, L9
Levison, H. F. & Duncan, M. J. 1994, Icarus, 108, 18
Lutz, B. L., Womack, M., & Wagner, R. M. 1993, ApJ, 407, 402
Luu, J. X. & Jewitt, D. C. 1990, AJ, 100, 913
Luu, J. X., Jewitt, D. C., & Trujillo, C. 2000, ApJ, 531, L151
Magnani, L. & A'Hearn, M. F. 1986, ApJ, 302, 477
Marcialis, R. L. & Buratti, B. J. 1993, Icarus, 104, 234
Matthews, C. 1995, in Astronomical Society of the Pacific Conference Series, Vol. 74,
Progress in the Search for Extraterrestrial Life., ed. G. S. Shostak, 95
Mazzotta Epifani, E., Palumbo, P., Capria, M. T., Cremonese, G., Fulle, M., & Colangeli,
L. 2007, MNRAS, 381, 713
McKay, A. J., Chanover, N. J., Morgenthaler, J. P., Cochran, A. L., Harris, W. M., &
Russo, N. D. 2012, Icarus, 220, 277
McKay, A. J., Cochran, A. L., DiSanti, M. A., Villanueva, G., Russo, N. D., Vervack, R. J.,
Morgenthaler, J. P., Harris, W. M., & Chanover, N. J. 2015, Icarus, 250, 504
Meech, K. J. & Belton, M. J. S. 1990, Astron. J., 100, 1323
Meech, K. J., Belton, M. J. S., Mueller, B. E. A., Dicksion, M. W., & Li, H. R. 1993, AJ,
106, 1222
Meech, K. J. & Jewitt, D. 1987, Nature, 328, 506
– 55 –
Miles, R. 2016, Icarus, 272, 356
Moreno, F. 2009, ApJS, 183, 33
Mumma, M. J., Weissman, P. R., & Stern, S. A. 1993, in Protostars and Planets III, ed.
E. H. Levy & J. I. Lunine, 1177–1252
Ootsubo, T., Kawakita, H., Hamada, S., Kobayashi, H., Yamaguchi, M., Usui, F.,
Nakagawa, T., Ueno, M., Ishiguro, M., Sekiguchi, T., Watanabe, J.-i., Sakon, I.,
Shimonishi, T., & Onaka, T. 2012, ApJ, 752, 15
Ortiz, J. L., Duffard, R., Pinilla-Alonso, N., Alvarez-Candal, A., Santos-Sanz, P., Morales,
N., Fern´andez-Valenzuela, E., Licandro, J., Campo Bagatin, A., & Thirouin, A.
2015, A&A, 576, A18
Paganini, L., Mumma, M. J., Boehnhardt, H., DiSanti, M. A., Villanueva, G. L., Bonev,
B. P., Lippi, M., Kaufl, H. U., & Blake, G. A. 2013, Astroph. J., 766, 100
Parker, J. W., Stern, S. A., Festou, M. C., A'Hearn, M. F., & Weintraub, D. 1997, AJ, 113,
1899
Patashnick, H. 1974, Nature, 250, 313
Patzold, M., Andert, T., Hahn, M., Asmar, S. W., Barriot, J.-P., Bird, M. K., Hausler, B.,
Peter, K., Tellmann, S., Grun, E., Weissman, P. R., Sierks, H., Jorda, L., Gaskell,
R., Preusker, F., & Scholten, F. 2016, Nature, 530, 63
Pierce, D. M. & A'Hearn, M. F. 2010, ApJ, 718, 340
Prialnik, D. 1997a, ApJ, 478, L107
-. 1997b, Earth Moon and Planets, 77, 223
– 56 –
Prialnik, D. & Bar-Nun, A. 1990a, ApJ, 363, 274
-. 1990b, ApJ, 363, 274
Prialnik, D., Bar-Nun, A., & Podolak, M. 1987, ApJ, 319, 993
Prialnik, D., Benkhoff, J., & Podolak, M. Modeling the structure and activity of comet
nuclei, ed. G. W. Kronk, 359–387
Prialnik, D., Brosch, N., & Ianovici, D. 1995, MNRAS, 276, 1148
Prialnik, D., Egozi, U., Bar-Nun, A., Podolak, M., & Mekler, Y. 1992, in Bulletin of
the American Astronomical Society, Vol. 24, AAS/Division for Planetary Sciences
Meeting Abstracts #24, 1019
Prialnik, D. & Podolak, M. 1995, Icarus, 117, 420
-. 1999, Space Sci. Rev., 90, 169
Prialnik, D., Sarid, G., Rosenberg, E. D., & Merk, R. 2008, Space Sci. Rev., 138, 147
Protopapa, S., Sunshine, J. M., Feaga, L. M., Kelley, M. S. P., A'Hearn, M. F., Farnham,
T. L., Groussin, O., Besse, S., Merlin, F., & Li, J.-Y. 2014, Icarus, 238, 191
Raghuram, S. & Bhardwaj, A. 2014, A&A, 566, A134
Rauer, H., Biver, N., Crovisier, J., Bockel´ee-Morvan, D., Colom, P., Despois, D., Ip, W.-H.,
Jorda, L., Lellouch, E., Paubert, G., & Thomas, N. 1997, Planet. Space Sci., 45, 799
Rauer, H., Helbert, J., Arpigny, C., Benkhoff, J., Bockel´ee-Morvan, D., Boehnhardt, H.,
Colas, F., Crovisier, J., Hainaut, O., Jorda, L., Kueppers, M., Manfroid, J., &
Thomas, N. 2003, A&A, 397, 1109
Reach, W. T., Kelley, M. S., & Vaubaillon, J. 2013a, Icarus, 226, 777
– 57 –
-. 2013b, Icarus, 226, 777
Rettig, T. W., Tegler, S. C., Pasto, D. J., & Mumma, M. J. 1992, ApJ, 398, 293
Roemer, E. 1958, PASP, 70, 272
-. 1962, PASP, 74, 351
Romon-Martin, J., Delahodde, C., Barucci, M. A., de Bergh, C., & Peixinho, N. 2003,
A&A, 400, 369
Rousselot, P. 2008, A&A, 480, 543
Rubin, M., Altwegg, K., Balsiger, H., Bar-Nun, A., Berthelier, J.-J., Bieler, A., Bochsler,
P., Briois, C., Calmonte, U., Combi, M., De Keyser, J., Dhooghe, F., Eberhardt,
P., Fiethe, B., Fuselier, S. A., Gasc, S., Gombosi, T. I., Hansen, K. C., Hassig, M.,
Jackel, A., Kopp, E., Korth, A., Le Roy, L., Mall, U., Marty, B., Mousis, O., Owen,
T., R`eme, H., S´emon, T., Tzou, C.-Y., Waite, J. H., & Wurz, P. 2015, Science, 348,
232
Ruprecht, J. D., Bosh, A. S., Person, M. J., Bianco, F. B., Fulton, B. J., Gulbis, A. A. S.,
Bus, S. J., & Zangari, A. M. 2015, Icarus, 252, 271
Sarid, G., Prialnik, D., Meech, K. J., Pittichov´a, J., & Farnham, T. L. 2005, PASP, 117,
796
Schambeau, C. A., Fern´andez, Y. R., Lisse, C. M., Samarasinha, N., & Woodney, L. M.
2015, Icarus, 260, 60
Schmitt, B., Espinasse, S., Grim, R. J. A., Greenberg, J. M., & Klinger, J. 1989, in ESA
Special Publication, Vol. 302, Physics and Mechanics of Cometary Materials, ed.
J. J. Hunt & T. D. Guyenne
– 58 –
Schulz, R. 2002, A&A Rev., 11, 1
Sekanina, Z. 1996, A&A, 314, 957
-. 1997, Earth Moon and Planets, 77, 147
Sekanina, Z., Chodas, P. W., & Yeomans, D. K. 1994, A&A, 289, 607
Sekanina, Z., Larson, S. M., Hainaut, O., Smette, A., & West, R. M. 1992, A&A, 263, 367
Senay, M. C. & Jewitt, D. 1994, Nature, 371, 229
Stansberry, J. A., Van Cleve, J., Reach, W. T., Cruikshank, D. P., Emery, J. P., Fernandez,
Y. R., Meadows, V. S., Su, K. Y. L., Misselt, K., Rieke, G. H., Young, E. T., Werner,
M. W., Engelbracht, C. W., Gordon, K. D., Hines, D. C., Kelly, D. M., Morrison,
J. E., & Muzerolle, J. 2004, Astroph. J. Supp., 154, 463
Stern, S. A. 1989, PASP, 101, 126
-. 1995, AJ, 110, 856
Stern, S. A., Cunningham, N. J., & Schindhelm, E. 2014, AJ, 147, 102
Strazzulla, G., Pirronello, V., & Foti, G. 1983, A&A, 123, 93
Tholen, D. J., Hartmann, W. K., Cruikshank, D. P., Lilly, S., Bowell, E., & Hewitt, A.
1988, IAU Circ., 4554
Thomas, N., Sierks, H., Barbieri, C., Lamy, P. L., Rodrigo, R., Rickman, H., Koschny,
D., Keller, H. U., Agarwal, J., A'Hearn, M. F., Angrilli, F., Auger, A.-T., Barucci,
M. A., Bertaux, J.-L., Bertini, I., Besse, S., Bodewits, D., Cremonese, G., Da Deppo,
V., Davidsson, B., De Cecco, M., Debei, S., El-Maarry, M. R., Ferri, F., Fornasier,
S., Fulle, M., Giacomini, L., Groussin, O., Gutierrez, P. J., Guttler, C., Hviid, S. F.,
– 59 –
Ip, W.-H., Jorda, L., Knollenberg, J., Kramm, J.-R., Kuhrt, E., Kuppers, M., La
Forgia, F., Lara, L. M., Lazzarin, M., Moreno, J. J. L., Magrin, S., Marchi, S.,
Marzari, F., Massironi, M., Michalik, H., Moissl, R., Mottola, S., Naletto, G., Oklay,
N., Pajola, M., Pommerol, A., Preusker, F., Sabau, L., Scholten, F., Snodgrass, C.,
Tubiana, C., Vincent, J.-B., & Wenzel, K.-P. 2015, Science, 347, aaa0440
Trigo-Rodr´ıguez, J. M., Garc´ıa-Hern´andez, D. A., S´anchez, A., Lacruz, J., Davidsson,
B. J. R., Rodr´ıguez, D., Pastor, S., & de Los Reyes, J. A. 2010, MNRAS, 409, 1682
Weaver, H. A., Feldman, P. D., A'Hearn, M. F., & Arpigny, C. 1997, Science, 275, 1900
Weaver, H. A. & Lamy, P. L. 1997, Earth Moon and Planets, 79, 17
Weiler, M., Rauer, H., Knollenberg, J., Jorda, L., & Helbert, J. 2003, A&A, 403, 313
Weissman, P. R. 1999, Space Sci. Rev., 90, 301
West, R. M. 1991, A&A, 241, 635
Whipple, F. L. 1980, AJ, 85, 305
Whitney, C. 1955, ApJ, 122, 190
Womack, M., Festou, M. C., & Stern, S. A. 1997, AJ, 114, 2789
Womack, M. & Stern, S. A. 1999, Solar System Research, 33, 187
Womack, M., Wyckoff, S., & Ziurys, L. M. 1992, ApJ, 401, 728
Woodney, L. M., A'Hearn, M. F., McMullin, J., & Samarasinha, N. 1997, Earth Moon and
Planets, 78, 69
Wyckoff, S., Wagner, R. M., Wehinger, P. A., Schleicher, D. G., & Festou, M. C. 1985,
Nature, 316, 241
– 60 –
Yamamoto, T. 1985, A&A, 142, 31
This manuscript was prepared with the AAS LATEX macros v5.2.
– 61 –
Table 1: Sublimation Temperatures of Cometary Species
Species Temperaturea (K)
N2
CO
CH4
H2S
C2H2
H2CO
CO2
HC3N
NH3
CS2
SO2
CH3CN
HCN
CH3OH
H2O
22
25
31
57
57
64
72
74
78
78
83
91
95
99
152
a(Yamamoto 1985; Sekanina 1996)
– 62 –
Table 2: Representative Gaseous Production Rates in Three Comets Beyond 4 AU
Comet
Molecule
r (AU) Q (number/s) Reference
29P
29P
29P
29P
29P
29P
29P
Chiron
Chiron
CN
CO
H2O
CO2
HCN
CH3OH
CH4
CN
CO
Chiron
H2O (via OH)
Chiron
H2S
Chiron
H2CO
Hale-Bopp CN
Hale-Bopp CN
Hale-Bopp CO
Hale-Bopp H2O
Hale-Bopp CO2
Hale-Bopp HCN
Hale-Bopp CH3OH
Hale-Bopp H2CO
Hale-Bopp C3
5.8
8.0 × 1024
(Cochran & Cochran 1991)
5.7 - 6.2
1-7 × 1028
variousa
6.2
6.2
6.3
6.3
6.3
6.3 × 1027
(Ootsubo et al. 2012)
<3.5e26
(Ootsubo et al. 2012)
< 1.4 × 1027
(Paganini et al. 2013)
<9.3 × 1027
(Paganini et al. 2013)
<1.3 × 1027
(Paganini et al. 2013)
11.3
3.7 × 1025
(Bus et al. 1991)
8.5
8.5
8.5
8.5
6.8
9.8
6.8
4.8
4.6
4.8
5.0
4.1
7.0
1.3 × 1028
(Womack & Stern 1999), recalculated
<4 × 1030
(Parker et al. 1997)
<2.1 × 1027
(Rauer et al. 1997)
<4.2 × 1026
(Rauer et al. 1997)
6 × 1025
(Fitzsimmons & Cartwright 1996)
0.9 × 1025
(Rauer et al. 2003)
1.8 × 1028
(Jewitt et al. 1996)
2.0 × 1028
(Weaver et al. 1997)
1.3 × 1028
(Crovisier 1997)
2 × 1026
(Jewitt et al. 1996)
3.3 × 1027
(Womack et al. 1997)
5.8 × 1026
(Biver et al. 1999)
≤ 0.6 × 1025
(Rauer et al. 2003)
aThe range of CO production rates typically observed is listed, taken from Senay & Jewitt (1994);
Crovisier et al. (1995); Festou et al. (2001); Gunnarsson et al. (2008); Paganini et al. (2013)
– 63 –
Table 3: CN/CO production rate ratios in three distant comets
Comet
r (AU) Q(CN)/Q(CO) Reference
Chiron
11, 8.5
0.0025
Bus et al. (1991); Womack & Stern (1999)
29P
5.8
Hale-Bopp 6.8
Hale-Bopp 6.0
0.0003
0.0033
0.0040
Cochran & Cochran (1991); Gunnarsson et al. (2008)a
Fitzsimmons & Cartwright (1996); Jewitt et al. (1996)
Rauer et al. (1997); Biver et al. (2002)
aThis ratio for 29P may typically be significantly lower, as CN is rarely observed in this comet.
– 64 –
Table 4: Ratios of CO2/CO production rates in distant comets
Comet
r (AU) Q(CO2)/Q(CO) Reference
29p
6.2
<0.0125
Ootsubo et al. (2012)
Hale-Bopp 4.6
Hale-Bopp 4.9
Hale-Bopp 4.0
Hale-Bopp 3.89
0.10
0.19
0.55
0.10
Crovisier (1997)
Crovisier (1997)
Weaver et al. (1997)
Crovisier (1997)
– 65 –
Table 5: CO Production Rates in Comet 29P
Date
r (AU) Qa ( × 1028 molecules s−1)
25-Feb-16
5.96
5.3 ± 1.3
26-Feb-16
5.96
3.8 ± 0.9
28-Feb-16
5.96
5.1 ± 1.3
29-Feb-16
5.96
2.5 ± 0.6
1-Mar-16
5.96
2.8 ± 0.7
3-Mar-16
5.96
2.6 ± 0.7
21-Mar-16
5.95
3.2 ± 0.8
28-Mar-16
5.95
2.6 ± 0.7
31-Mar-16
5.95
1.0 ± 0.2
9-Apr-16
5.95
0.9 ± 0.2
15-Apr-16
5.95
0.8 ± 0.2
24-Apr-16
5.94
1.0 ± 0.2
29-May-16
5.93
1.0 ± 0.3
aProduction rates derived from observations of the CO J=2-1 line using the Arizona Radio Observatory 10-m
Submillimeter Telescope (SMT). Full details in Wierzchos et al., in preparation.
– 66 –
Comet Hale-Bopp
Heliocentric Distance ~ 5 AU
CO
CH3OH
-6
-4
-2
0
2
4
6
8
Offset Velocity (km/s)
)
K
(
*
R
T
0.1
0
0.05
0.04
0.03
0.02
0.01
0
-0.01
-0.02
-8
Fig. 1.- Spectra of CO and CH3OH emission from comet C/1995 O1 Hale-Bopp at 5
AU, pre-perihelion, obtained with the NRAO 12-m telescope. The CO J=2-1 spectrum was
obtained at 230 GHz on 1996 March 31 with 49 kHz/channel resolution, which corresponds
to 0.06 km/s/channel; the methanol emission was obtained at 145 GHz on 1996 Apr 16
with 195 kHz/channel resolution, which corresponds to 0.40 km/s/channel. The dotted line
indicates the comet's rest frame velocity. As the figure shows, the CO emission is blue-shifted
from the ephemeris velocity of the comet and has a very narrow line-width. In contrast, the
methanol emission has no measurable shift from the comet's velocity and has a broader
line-width. This difference can be explained if CO emission originates from within the cold
comet nucleus and the methanol escapes from hot grains in the coma.
– 67 –
CO Production Rates in 3 Distant Comets
1029
1028
1027
)
s
/
s
e
l
u
c
e
l
o
m
(
)
O
C
Q
(
Hale-Bopp: Womack et al. 1997, re-calculated
29P: Graham & Womack 1995 recalculated, Wierzchos et al., in prep.
Chiron: Womack & Stern 1999, recalculated
Chiron: Drahus et al. 2016, upper limit
Hale-Bopp: Gunnarsson et al. 2003, isotropic source
Q(CO)=3.5x1029r-2 for Hale-Bopp
4
5
6
Heliocentric distance (AU)
7
8
9
10
10
11
Fig. 2.- CO production rates from three distant comets are plotted against heliocentric
distance. The observations of Hale-Bopp and 29P are consistent with a nominal production
rate of Q(CO)=3.5 × 1029r−2 superimposed with sporadic outbursts.
– 68 –
Comet 29P/Schwassmann-Wachmann 1
Heliocentric Distance ~ 6 AU
CO J=2-1
0.1
0.05
0
)
K
(
*
R
T
-4
-2
0
Offset Velocity (km/s)
2
4
Fig. 3.- Spectrum of the CO J=2-1 rotational line in comet 29P/Schwassmann-Wachmann
1 obtained from 1995 Dec 9–11 UT with the NRAO 12-m telescope (Graham & Womack
1995). The spectral resolution was 100 kHz, which corresponds to 0.13 km/s per channel.
The dotted line indicates the comet's rest frame velocity. Two narrow velocity components
are seen in this spectrum and in many other observations of this comet.
– 69 –
Specific Production Rates for CO in Distant Comets
)
2
m
k
/
s
/
s
e
l
u
c
e
l
o
m
(
2
D
/
)
O
C
Q
(
1027
1026
1025
1024
1023
1022
4
4
5
5
Hale-Bopp
10x enriched
50x enriched
10x depleted
50x depleted
29P
Chiron
Chiron upper limit
8
8
9
9
10
10
6
6
7
7
Heliocentric distance (AU)
Heliocentric distance (AU)
Fig. 4.- Specific production rates, Q(CO)/D2, for distant comets. The solid line is the
specific production rate derived for Hale-Bopp, assuming Q(CO)=3.5 × 1029r−2, and diam-
eter DHale−Bopp=60km. Values for Chiron and 29P are plotted assuming production rates
from Figure 2 and DChiron=218km and D29P =60 km. This plot shows that, when adjusted
for surface area and heliocentric distance, 29P (in non-outbursting mode) produces CO at
approximately the same rate as, and up to ten times larger than, Hale-Bopp. In contrast,
Chiron produces at least 5-15 times less CO than Hale-Bopp. This may indicate that 29P
entered its current orbit more recently than many models predict, and that Chiron is signif-
icantly depleted in CO over its initial chemical composition.
|
1306.0610 | 2 | 1306 | 2013-08-13T18:36:03 | A Combined VLT and Gemini Study of the Atmosphere of the Directly-Imaged Planet, beta Pictoris b | [
"astro-ph.EP",
"astro-ph.IM",
"astro-ph.SR"
] | We analyze new/archival VLT/NaCo and Gemini/NICI high-contrast imaging of the young, self-luminous planet $\beta$ Pictoris b in seven near-to-mid IR photometric filters, using advanced image processing methods to achieve high signal-to-noise, high precision measurements. While $\beta$ Pic b's near-IR colors mimick that of a standard, cloudy early-to-mid L dwarf, it is overluminous in the mid-infrared compared to the field L/T dwarf sequence. Few substellar/planet-mass objects -- i.e. $\kappa$ And b and 1RXJ 1609B -- match $\beta$ Pic b's $JHK_{s}L^\prime$ photometry, and its 3.1 $\mu m$ and 5 $\mu m$ photometry are particularly difficult to reproduce. Atmosphere models adopting cloud prescriptions and large ($\sim$ 60 $\mu m$) dust grains fail to reproduce the $\beta$ Pic b spectrum. However, models incorporating thick clouds similar to those found for HR 8799 bcde but also with small (a few microns) modal particle sizes yield fits consistent with the data within uncertainties. Assuming solar abundance models, thick clouds, and small dust particles ($<a>$ = 4 $\mu m$) we derive atmosphere parameters of log(g) = 3.8 $\pm$ 0.2 and $T_{eff}$ = 1575--1650 $K$, an inferred mass of 7$^{+4}_{-3}$ $M_{J}$, and a luminosity of log(L/L$_{\odot}$) $\sim$ -3.80 $\pm$ 0.02. The best-estimated planet radius, $\approx$ 1.65 $\pm$ 0.06 $R_{J}$, is near the upper end of allowable planet radii for hot-start models given the host star's age and likely reflects challenges with constructing accurate atmospheric models. Alternatively, these radii are comfortably consistent with hot-start model predictions if $\beta$ Pic b is younger than $\approx$ 7 Myr, consistent with a late formation, well after its host star's birth $\sim$ 12$^{+8}_{-4}$ Myr ago. | astro-ph.EP | astro-ph | Draft version July 22, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
A COMBINED VLT AND GEMINI STUDY OF THE ATMOSPHERE OF THE DIRECTLY-IMAGED PLANET,
β PICTORIS b
Thayne Currie1, Adam Burrows2, Nikku Madhusudhan3, Misato Fukagawa4, Julien H. Girard5, Rebekah
Dawson6, Ruth Murray-Clay6, Scott Kenyon6, Marc Kuchner7, Soko Matsumura8, Ray Jayawardhana1, John
Chambers9, Ben Bromley10
3
1
0
2
g
u
A
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
0
1
6
0
.
6
0
3
1
:
v
i
X
r
a
Draft version July 22, 2018
ABSTRACT
We analyze new/archival VLT/NaCo and Gemini/NICI high-contrast imaging of the young, self-
luminous planet β Pictoris b in seven near-to-mid IR photometric filters, using advanced image pro-
cessing methods to achieve high signal-to-noise, high precision measurements. While β Pic b's near-IR
colors mimick that of a standard, cloudy early-to-mid L dwarf, it is overluminous in the mid-infrared
compared to the field L/T dwarf sequence. Few substellar/planet-mass objects -- i.e. κ And b and
1RXJ 1609B -- match β Pic b's JHKsL′ photometry, and its 3.1 µm and 5 µm photometry are
particularly difficult to reproduce. Atmosphere models adopting cloud prescriptions and large (∼ 60
µm) dust grains fail to reproduce the β Pic b spectrum. However, models incorporating thick clouds
similar to those found for HR 8799 bcde but also with small (a few microns) modal particle sizes yield
fits consistent with the data within uncertainties. Assuming solar abundance models, thick clouds,
and small dust particles (< a > = 4 µm) we derive atmosphere parameters of log(g) = 3.8 ± 0.2 and
Tef f = 1575 -- 1650 K, an inferred mass of 7+4
−3 MJ , and a luminosity of log(L/L⊙) ∼ -3.80 ± 0.02.
The best-estimated planet radius, ≈ 1.65 ± 0.06 RJ , is near the upper end of allowable planet radii
for hot-start models given the host star's age and likely reflects challenges with constructing accurate
atmospheric models. Alternatively, these radii are comfortably consistent with hot-start model pre-
dictions if β Pic b is younger than ≈ 7 Myr, consistent with a late formation, well after its host star's
birth ∼ 12+8
Subject headings: planetary systems, stars: early-type, stars: individual: β Pictoris
−4 Myr ago.
1. INTRODUCTION
The method of detecting extrasolar planets by di-
rect imaging, even in its current early stage, fills in
an important gap in our knowledge of the diversity of
planetary systems around nearby stars. Direct imag-
ing searches with the best conventional AO systems (e.g.
Keck/NIRC2, VLT/NaCo, Subaru/HiCIAO) are sensi-
tive to very massive planets (M & 5 -- 10 MJ ) at wide
separation (a ∼ 10-30 AU to 100 AU ) and young ages (t
. 100 Myr), which are not detectable by the radial ve-
locity and transit methods (e.g. Lafreni`ere et al. 2007a;
Vigan et al. 2012; Rameau et al. 2013; Galicher et al.
2013). Planets with these masses and orbital sepa-
rations pose a stiff challenge to planet formation the-
ories (e.g. Kratter et al. 2010; Rafikov 2011). Young
self-luminous directly-imageable planets provide a criti-
cal probe of planet atmospheric evolution (Fortney et al.
2008; Currie et al. 2011a; Spiegel and Burrows 2012;
Konopacky et al. 2013).
1 Department of Astronomy and Astrophysics, University of
Toronto
2 Department of Astrophysical Sciences, Princeton University
3 Department of Astronomy, Yale University
4 Osaka University
5 European Southern Observatory
6 Harvard-Smithsonian Center for Astrophysics
7 NASA-Goddard Space Flight Center
8 Department of Astronomy, University of Maryland-College
Park
9 Department of Terrestrial Magnetism, Carnegie Institution
of Washington
10 Department of Physics, University of Utah
The directly-imaged planet around the nearby star β
Pictoris (β Pictoris b) is a particularly clear, crucial
test for understanding the formation and atmospheric
evolution of gas giant planets (Lagrange et al. 2009a,
2010). At 12+8
−4 Myr old (Zuckerman et al. 2001), the
β Pictoris system provides a way to probe planet at-
mospheric properties only ≈ 5 -- 10 Myr after the disks
from which planets form dissipate (≈ 3 -- 10 Myr, e.g.
Pascucci et al. 2006; Currie et al. 2009).
Similar to
the case for the HR 8799 planets (Marois et al. 2010a;
Fabrycky and Murray-Clay 2010; Currie et al. 2011a;
Sudol and Haghighipour 2012), β Pic b's mass can
be constrained without depending on highly-uncertain
planet cooling models: in this case, RV-derived dynam-
ical mass upper limits when coupled with the range of
plausible orbits (a ∼ 8 -- 10 AU) imply masses less than
∼ 10 -- 15 MJ (Lagrange et al. 2012a; Currie et al. 2011b;
Chauvin et al. 2012; Bonnefoy et al. 2013), a mass range
consistent with estimates derived from the planet's inter-
action with the secondary disk (Lagrange et al. 2009a;
Dawson et al. 2011).
Furthermore, while other likely/candidate planets
such as Fomalhaut b and LkCa 15 b are proba-
bly made detectable by circumplanetary emission in
some poorly constrained geometery (Currie et al. 2012a;
Kraus and Ireland 2012), β Pic b's emission appears to
be consistent with that from a self-luminous planet's
atmosphere (Currie et al. 2011b; Bonnefoy et al. 2013).
Other objects of comparable mass appear to have formed
more like low-mass binary companions. Thus, com-
bined with the planets HR 8799 bcde, β Pic b pro-
2
vides a crucial reference point with which to interpret
the properties of many soon-to-be imaged planets with
upcoming extreme AO systems like GP I, SCExAO, and
SP HERE (Macintosh et al. 2008; Martinache et al.
2009; Beuzit et al. 2008).
However, investigations into β Pic b's atmosphere are
still in an early stage compared to those for the at-
mospheres of the HR 8799 planets and other very low-
mass, young substellar objects (e.g. Currie et al. 2011a;
Skemer et al. 2011; Konopacky et al. 2013; Bailey et al.
2013). Of the current published photometry, only
Ks (2.18 µm) and L′ (3.78 µm) have photometric er-
rors smaller than ∼ 0.1 mag (Bonnefoy et al. 2011;
Currie et al. 2011b). Other high SNR detections such
as at M ′ were obtained without reliable flux calibra-
tion (Currie et al. 2011b) or with additional, large pho-
tometric uncertainties due to processing (Bonnefoy et al.
2013). As a result, the best-fit models admit a wide range
of temperatures, surface gravities, and cloud structures
(e.g. Currie et al. 2011b). Thus, new higher signal-to-
noise/precision and flux-calibrated photometry at 1 -- 5
µm should provide a clearer picture of the clouds, chem-
istry, temperature, and gravity of β Pic b. Moreover, new
near-to-mid IR data may identify distinguishing char-
acteristics of β Pic b's atmosphere, much like clouds
and non-equilibrium carbon chemistry for HR 8799 bcde
(Currie et al. 2011a; Galicher et al. 2011; Skemer et al.
2012; Konopacky et al. 2013).
In this study, we present new 1.5 -- 5 µm observations
for β Pic b obtained with N aCo on the Very Large Tele-
scope and N ICI on Gemini-South. We extract the first
detection at the 3.09 µm water-ice filter; the first high
signal-noise, well calibrated H, [4.05], and M ′ detections;
and higher signal-to-noise detections at Ks and L′ (2.18
and 3.8 µm). To our new data, we add rereduced β Pic
data obtained in J (1.25 µm) and H (1.65 µm) bands
and first presented in Bonnefoy et al. (2013), recovering
β Pic b at a slightly higher signal-to-noise and deriving
its photometry with smaller errors.
We compare the colors derived from broadband pho-
tometry to that for field substellar objects with a range
of spectral types to assess whether β Pic b's col-
ors appear anomalous/redder than the field sequence
like those for planets around HR 8799 and κ And;
planet-mass companions like 2M 1207 B, GSC 06214 B,
and 1RXJ 1609 B (Chauvin et al. 2004; Ireland et al.
2011; Lafreni´ere et al. 2008); and other substellar ob-
jects like Luhman 16B (Luhman et al. 2013). We use
atmosphere modeling to constrain the range of tem-
peratures, surface gravities, and cloud structures plau-
sible for the planet. While previous studies have
shown the importance of clouds and non-equilibrium
carbon chemistry in fitting the spectra/photometry of
directly-imaged planets (Bowler et al. 2010; Currie et al.
2011a; Madhusudhan et al. 2011; Galicher et al. 2011;
Skemer et al. 2012; Konopacky et al. 2013), here the as-
sumed sizes of dust particles entrained in the clouds plays
a critical role.
2. OBSERVATIONS AND DATA REDUCTION
2.1. VLT/NaCo Data and Basic Processing
We observed β Pictoris under photometric condi-
tions on 14 December to 17 December 2012 with
the NAOS-CONICA instrument (NaCo; Rousset et al.
2003) on the Very Large Telescope UT4/Yepun at
Paranal Observatory (Program ID 090.C-0396). All data
were taken in pupil-tracking/angular differential imaging
(Marois et al. 2006) and data cube mode. Table 1 sum-
marizes the basic properties of these observations. Our
full complement of data during the run includes imag-
ing at 1.04 µm, 2.12 µm, Ks/2.18 µm, 2.32 µm, 3.74
µm, L′/3.78 µm, Br-α/4.05 µm, and M ′. Here, we fo-
cus only on the L′, [4.05], and M ′ data, deferring the
rest to a later study. Each observation was centered on
β Pictoris's transit for a total field rotation of ∼ 50 -- 70
degrees and a total observing times ranging between ∼
30 minutes and 59 minutes.
To these new observations, we rereduce J-band and H-
band data first presented in Bonnefoy et al. (2013) and
taken on 16 December 2011 and 11 January 2012, re-
spectively. The saturated J band science images are
bracketed by two sequences of unsaturated images ob-
tained in neutral density filter for flux calibration. While
there were additional frames taken but not analyzed in
Bonnefoy et al., we found these to be of significantly
poorer quality and thus do not consider them here. In
total, the J-band data we consider covers 40 minutes of
integration time and ∼ 23o of field rotation. The H-band
data cover ∼ 92 minutes of integration time and ∼ 36o
of field rotation.
Basic NaCo image processing steps were performed as
in Currie et al. (2010, 2011b). The thermal IR data at
L′ and [4.05] (M ′) were obtained in a dither pattern with
offsets every 2 (1) images to remove the sky background.
As all data were obtained in data cube mode, we in-
creased our PSF quality by realigning each individual
exposure in the cube to a common center position and
clipping out frames with low encircled energy (i.e. those
with a core/halo ratio < max(core/halo) - 3×σ(core-to-
halo ratio)).
2.2. Gemini/NICI Data and Basic Processing
We obtained Gemini imaging for β Pic b using the
Near-Infrared Coronagraphic Imager (NICI) on 23 De-
cember 2012 and 26 December 2012 in the H2O filter
(λo = 3.09 µm) and 9 January 2013 in the H and Ks
filters (dual-channel imaging), both under photometric
conditions (Program GS-2012B-Q-40). These observa-
tions were also executed in angular differential imaging
mode. For the H2O data, we dithered each 38 s expo-
sure for sky subtraction for a total of ∼ 38 minutes of
integration time over a field rotation of ∼ 30 degrees.
For the H/Ks data, we placed the star behind the r =
0.′′22 partially-transmissive coronagraphic mask to sup-
press the stellar halo. Here, we took shorter exposures
of β Pic (tint ∼ 11.4 s) to better identify and filter out
frames with bad AO correction. Our observing sequence
consists of ∼ 22 minutes of usable data centered on tran-
sit with a field rotation of ∼ 41 degrees.
Basic image processing follows steps described above
for NaCo data. The PSF halo was saturated out to r ∼
0.′′32 -- 0.′′36 in H during most of the observations and our
sequence suffered periodic seeing bubbles that saturated
the halo out to angular separations overlapping with the
β Pic b PSF. Thus, we focus on reducing only those
H-band frames with less severe halo saturation (rsat <
0.′′36). The Ks observations, obtained at a higher Strehl
ratio, never suffered halo saturation. The first of the
two H20 sets, suffered from severe periodic seeing bub-
bles and thus generally poor AO performance. We iden-
tify and remove from analysis frames whose halo flux
exceeded the Fmin+3σ, where Fmin is the minimum flux
within an aperture covering β Pic b and σ is the disper-
sion in this flux: about 10-25% of the frames, depending
on the data set in question.
2.3. PSF Subtraction
To remove the noisy stellar halo and reveal β Pic b, we
process the data with our "adaptive" LOCI (A-LOCI)
pipeline (Currie et al. 2012a,b, T. Currie 2013 in prep.).
This approach adopts "locally optimized combination
of images" (LOCI) formalism (Lafreni`ere et al. 2007b),
where we perform PSF subtraction in small annular re-
gions (the "subtraction zone") at a time over each im-
age. Previously-described A-LOCI components we use
here include "subtraction zone centering" (Currie et al.
2012b); "speckle filtering" to identify and remove images
with noise structure poorly correlated with that from the
science image we are wanting to subtract (Currie et al.
2012b); a moving pixel mask to increase point source
throughput and normalize it as a function of azimuthal
angle (Currie et al. 2012a). We do not consider a PSF
reference library (Currie et al. 2012a) since β Pictoris is
our only target.
Into A-LOCI as recently utilized in Currie et al.
(2012a), we incorporate a component different from but
complementary to our "speckle filtering", using singular
value decomposition (SVD) to limit the number of images
used in a given annular region (i.e. for a given optimiza-
tion zone) to construct and subtract a reference. Briefly,
in the (A-)LOCI formalism a matrix inversion yields the
set of coefficients ck applied to each image making up the
reference "image": ck = A−1b. Here, A is the covariance
matrix and b is a column matrix defined from i pixels
in the "optimization zones" of the j-th reference image
i OT
section Oj and the science image, OT : bj = Pi
i
(see Lafreni`ere et al. 2007b). In the previous versions of
our codes, we use a simple double-precision matrix inver-
sion to invert the covariance matrix and then solve for
ck after multiplying by b.
Oj
In this work, we instead use SVD to rewrite A as
UΣVT such that A−1 = VΣ−1UT , where the T super-
script stands for the transpose of the matrix. Prior to
inversion, we truncate the number of singular values at a
predefined cutoff, svdlim. This eigenvalue truncation is
very similar to and functions the same as the truncation
of principle components, Npca, in the Karhunen-Loeve
image projection (KLIP) (Soummer et al. 2012) and
has been successfully incorporated before (Marois et al.
2010b). We found that both speckle filtering and SVD
truncation within our formalism can yield significant con-
trast gains over LOCI and KLIP/Principle Component
Analysis (PCA), although in this study at the angular
separation of β Pic b (≈ 0.′′45) the gains over LOCI are
typically about a factor of 1.5, albeit with substantially
higher throughput11.
11 Recently, Amara and Quanz (2012) claimed a contrast gain
of ∼ 5× over LOCI using PCA. However, optimal set-ups even
within a given formalism like LOCI or PCA/KLIP are very dataset-
3
2.4. Planet Detections
Figures 1, 2, and 3 display reduced NaCo and NICI
images of β Pic. We detect β Pic b in all datasets (sum-
marized in Table 2). To compute the signal-to-noise ra-
tio (SNR) for β Pic b, we determine the dispersion, σ, in
pixel values of our final image convolved with a gaussian
along a ring with width of 1 FWHM at the same angu-
lar separation as β Pic b but excluding the planet (e.g.
Thalmann et al. 2009), and average the SNR/pixel over
the aperture area. For the Gemini-NICI H, Ks, and two
[3.1] datasets, the SNRs are thus 6.4, 11, 4.6, and 10, re-
spectively. For the J and H-band NaCo data previously
presented in Bonnefoy et al. (2013), we achieve SNR ∼
9 and SNR ∼ 30, respectively. Generally speaking, our
3.8 -- 5 µm NaCo data are deeper than the near-IR NaCo
and especially near-IR NICI data, where we detect β Pic
b at SNR = 40 in L′ and 22 at M ′, roughly a factor of
two higher than previously reported (Currie et al. 2011b;
Bonnefoy et al. 2013), gains due to β Pic b now being
at a wider projected separation (L′) or post-processing
and slightly better observing conditions (M ′). The high
SNR detections obtained with NaCo also leverage on re-
cent engineering upgrades that substantially improved
the instrument's image quality and the stability of its
PSF (Girard et al. 2012).
The optimal A-LOCI algorithm parameters vary sig-
nificantly from dataset to dataset. The rotation gap
(∆PA in units of the image full-width half maximum)
criterion used to produce most of the images is δ ∼ 0.6 --
0.65, although it is significantly larger for the J and H
data sets (δ = 0.75 -- 0.95). Generally speaking, the opti-
mization areas we use NA are significantly smaller (NA
= 50-150) than those typically adopted (i.e. NA=300;
Lafreni`ere et al. 2007b). We speculate that the pixel
masking component of A-LOCI drives the optimal NA
settings toward these smaller values since the planet flux
(ostensibly within the subtraction zone) no longer signif-
icantly biases the coefficient determinations to the point
of reducing the planet's SNR. Filtering parameters rcorr
and svdlim likewise vary wildly from rcorr = 0 and svdlim
= 2.5×10−7 at J to rcorr = 0.9 for the NICI H-band data
or svdlim = 2.5×10−2 for the M ′ NaCo data.
While the many algorithm free parameters make find-
ing an optimal combination difficult and computation-
ally expensive, our final image quality is nevertheless
extremely sensitive to some values, in particular svdlim
and rcorr. As a test, we explored other image process-
ing methods -- ADI-based classical PSF subtraction and
LOCI. While A-LOCI always yields deeper contrasts, we
easily detect β Pic b in the mid-IR NaCo data using any
method and only the poorer of the two [3.1] data sets re-
quires A-LOCI to yield a better than 4-σ detection (i.e.
where σdet = 1.0857/SNR = 0.27 mags). We will present
a detailed analysis of image processing methods and al-
gorithm parameters in an upcoming study (T. Currie,
2013 in prep.).
Adopting the pixel scales listed in Table 1, β Pic b is
specific (cf. Lafreni`ere et al. 2007b; Currie et al. 2012a,b). With
LOCI, we obtained roughly equivalent SNRs for β Pic b obtained
during the same observing run but on a night with poorer ob-
serving conditions (29 December 2009) than their test data set
(Currie et al. 2011b).
Implementing some A-LOCI filtering and
pixel masking yields SNR ≈ 30 -- 35.
4
detected at an angular separation of r ∼ 0.′′46 in each
data set. The position angle of β Pic b is consistent with
previously-listed values (PA ≈ 210o) and in between val-
ues for the main disk and the warp, intermediate be-
tween the results presented in Currie et al. (2011b) and
Lagrange et al. (2012b). While the NICI north position
angle on the detector is precisely known and determined
from facilty observations, we have not yet used our as-
trometric standard observations to derive the NaCo po-
sition angle offset, which changes every time NaCo is re-
moved from the telescope. To dissuade others from using
the poorly calibrated NaCo data and precisely calibrated
data (Lagrange et al. 2012b) together, we reserve a de-
tailed determination of β Pic b's astrometry and a study
of its orbit for a future study. We also detect the β Pic
debris disk in each new broadband data set and at [4.05]
(Figure 4). We will analyze its properties at a later time
as well.
2.5. Planet Photometry
To derive β Pic b photometry, we first measured its
brightness within an aperture roughly equal to the image
FWHM in each case, which was known since we either
had AO-corrected standard star observations (NICI H,
Ks, and [3.1]), unsaturated images of the primary as seen
through the coronagraphic mask (NICI Ks), unsaturated
neutral density filter observations (NaCo J, H, L′, and
M ′), or unsaturated images of the primary (NaCo L′ and
[4.05]). We assessed and corrected for planet throughput
losses due to processing by comparing the flux of syn-
thetic point sources within this aperture implanted into
registered images at the same angular separation as β
Pic b before and after processing. To derive β Pic b's
throughput and uncertainty in the throughput (σatten),
we repeat these measurements at 15 different position
angles and adopt the clipped mean of the throughput as
our throughput and standard deviation of this mean as
its uncertainty. The planet throughput ranges from 0.38
for the J-band data to 0.82 for the [4.05] data and 0.96
for the NICI H-band data, even with aggressive algo-
rithm parameters (i.e. δ ∼ 0.6), due to the throughput
gains yielded by our pixel masking and the SVD cutoff.
For photometric calibration, we followed several dif-
ferent approaches. For the NICI data, we used TYC
7594-1689-1 and HD 38921 as photometric standards.
We were only able to obtain photometric calibrations for
the first of the two [3.1] datasets. For all other data we
used the primary star, β Pic, for flux calibration adopt-
ing the measurements listed in Bonnefoy et al. (2013).
For the J and H NaCo data, we used images of the pri-
mary as viewed through the neutral density filter. For
the M ′ NaCo data, we obtained neutral density filter
observations and very short exposures. While the lat-
ter were close to saturation and were probably in the
non-linear regime, the implied photometry for β Pic was
consistent to within errors. The primary was unsatu-
rated in the [4.05]. Finally, for the L′ data, we took
8.372 ms unsaturated images of β Pic for flux calibra-
tion. In all cases, we again adopt the clipped mean of
individual measurements as our photometric calibration
uncertainty, σf luxcal. To compute the photometric un-
certainty for each data set, we considered the SNR of
our detection, the uncertainty in the planet throughput,
and the uncertainty in absolute flux calibration: σ =
qσ2
f luxcal.
det + σ2
atten + σ2
Table 2 reports our photometry and Table 3 lists sam-
ple error budgets for two NICI photometric measure-
ments and two NaCo measurements. The relative contri-
butions from each source of photometric uncertainty to
the total uncertainty are representative of our combined
data set. For the [3.09] data, residual speckle noise/sky
fluctuations greatly limit the planets SNR and thus σdet
is the primary source of photometric uncertainty. For the
Ks data, the intrinsic SNR and the two other sources of
photometric uncertainty contribute in a more equal pro-
portion. The L′ and M ′ data error budgets are charac-
teristic of most of our other data, where the photometric
uncertainty is primarily due to the absolute photomet-
ric calibration and throughput. With the exception of
the [3.09] NICI data, the intrinsic SNR of the detection
does not dominate the error budget. For the best-quality
(mid-IR NaCo) data, the throughput uncertainty was
small (≈ 5%) and was never any larger than 15% (J
band data) in any data set12.
In general, we find fair agreement with previously pub-
lished photometry, where our measurements are usually
consistent within photometric errors with those reported
previously (e.g. mH = 13.32 ± 0.14 and 13.25 ± 0.18 vs.
13.5 ± 0.2 in Bonnefoy et al. 2013). Our L′ photometry
is more consistent with Currie et al.'s measurement of
mL′=9.73 ± 0.06 than with that listed in Bonnefoy et al.
(2013) (mL′=9.5 ± 0.2), though it is nearly identical
to that derived for some β Pic b data sets listed in
Lagrange et al. (2010). Our [4.05] photometry implies
that β Pic b is ∼ 15-20% brighter there than previously
assumed (Quanz et al. 2010) and may have a slightly red
L′-[4.05] color. The major difference from previous stud-
ies, though, is that our photometric errors are consis-
tently much smaller. For example, the uncertainty in
the [4.05] photometry is reduced to 0.08 mag from 0.23
mag due both to higher SNR detections and lower un-
certainty in our derived photometry (e.g.
throughput
corrections). NICI photometry is also substantially less
uncertain than in Boccaletti et al. (2013) because β Pic
b is not occulted by the focal plane mask. These lower
uncertainties should allow more robust comparisons be-
tween β Pic b and other substellar objects and, from
modeling, more precise limits on the best-fitting planet
atmosphere properties.
3. EMPIRICAL COMPARISONS TO β PIC B
12 In principle, tuning the algorithm parameters to maximize
the SNR of β Pic b could introduce additional photometric un-
certainties if the planet is in significant residual speckle contami-
nation. In such a case, the algorithm parameters maximizing the
SNR could instead be the set that maximizes the residual speckle
contamination within the the planet aperture while minimizing it
elsewhere, especially as the pixel masking technique normalizes the
point source throughput but not the noise as a function of az-
imuthal angle. However, we do not find substantial differences in
the derived photometry if we adopt a default set of algorithm pa-
rameters. Furthermore, the parameters maximizing the SNR are
never the ones maximizing the planet throughput, and our tun-
ing is not just finding the parameter set making pixels within the
planet aperture 'noisiest'. Adopting slightly different parameters
from the 'optimized' case yields nearly identical photometry. More-
over, residual speckle contamination in most data sets is extremely
low, and for the mid-IR data the intrinsic SNR is limited by sky
background fluctuations in addition to speckles.
Our new data allows us to compare the spectral energy
distribution of β Pic b to that for the many field L/T-
type brown dwarfs as well that for directly-imaged low-
surface gravity, low-mass brown dwarf companions and
directly-imaged planets. Our goal here is to place β Pic
b within the general L/T type spectral sequence, iden-
tify departures from this sequence such as those seen for
low surface gravity objects like HR 8799 bcde, and iden-
tify the substellar object(s) with the best-matched SED.
Some bona fide directly-imaged planets like HR 8799
bcde and at least some of the lowest-mass brown dwarfs
like 2M 1207 B appear redder/cloudier than their field
dwarf counterparts at comparable temperatures (Tef f
≈ 900-1100 K). However, it is unclear whether hotter
imaged exoplanets appear different from their (already
cloudy) field L dwarf counterparts, and β Pic b provides
a test of any such differences. We will use our compar-
isons to the L/T dwarf sequence and the SEDs of other
substellar objects to inform our atmosphere model com-
parisons later to derive planet physical parameters (e.g.
Tef f and log(g)).
3.1. Infrared Colors of β Pic b
To compare the near-to-mid IR properties of β Pic
b with those for other cool, substellar objects, we
primarily use the sample of L/T dwarfs compiled by
Leggett et al. (2010), which include field dwarfs spec-
tral classes between ∼ M7 and T5, corresponding to
a range of temperatures between ∼ 2500 K and 700
K. To explore how the β Pic b SED compares to
those with other directly-imaged planets/planet candi-
dates and very low-mass brown dwarf companions within
this temperature range, we include objects listed in Ta-
ble 4. These include the directly-imaged planets around
HR 8799 (Marois et al. 2008, 2010a; Currie et al. 2011a)
and the directly-imaged planet candidate around κ And
(Carson et al. 2013). Additionally, we include high mass
ratio brown dwarf companions with masses less than
the deuterium-burning limit (∼ 13 -- 14 MJ ) and higher-
mass companions whose youth likely favors a lower
surface gravity than for field brown dwarfs, a differ-
ence that affect the objects' spectra (e.g. Luhman et al.
2007). Among these objects are 1RXJ 1609B, AB Pic B,
and Luhman 16 B (Lafreni´ere et al. 2008; Chauvin et al.
2005; Luhman et al. 2013). Table 5 compiles photometry
for all of these low surface gravity objects.
Figure 5 compares the IR colors of β Pic b (dark
blue diamonds) to those for field M dwarfs (small
black dots), field L0 -- L5 dwarfs (grey dots), field L5.1-
L9 dwarfs (asterisks), T dwarfs (small light-grey dots),
and planets/low-mass young brown dwarfs (light-blue
squares). The J-H/H-Ks colors for β Pic b appear
slightly blue in J-H and red in H-Ks compared to field
L0 -- L5 dwarfs, though the difference here is not as large
as was found in Bonnefoy et al. (2013). Other young
substellar objects appear to have similar near-IR colors,
in particular κ And b, GSC 06214 B, USco CTIO 108B,
2M 1207A, and Luhman 16 B, whose spectral types range
between M8 and T0.5.
The mid-IR colors of β Pic b (top-right and bottom
panels) show a more complicated situation. In J-Ks/Ks-
L′ and H-Ks/Ks-L′, β Pic b lies along the field L/T
dwarf locus with colors in between those for L0 -- L5 and
L5.1 -- L9 dwarfs, overlapping in color with κ And b, 1RXJ
5
1609B, GSC 06214B, HR 8799 d, and 2M 1207 B. Com-
pared to the few field L/T dwarfs from the Leggett et al.
sample with M ′ photometry, β Pic b appears rather red,
most similar in Ks-M ′ color to GSC 06214 B.
The color-magnitude diagram positions of β Pic b (Fig-
ure 6) better clarify how its near-to-mid SED compares to
the field L/T dwarf sequence and to very low-mass (and
gravity?) young substellar objects. In general, compared
to the field L dwarf sequence, β Pic b appears progres-
sively redder at mid-IR wavelengths. Similar to the case
for GSC 06214 B (Bailey et al. 2013), β Pic b appears
overluminous compared to the entire L/T dwarf sequence
in the mid-IR.
3.2. Comparisons to SEDs of Other Substellar Objects
To further explore how the SED of β Pic b agrees
with/departs from the field L/T dwarf sequence and
other young substellar objects, we first compare its pho-
tometry to spectra from the SPeX library (Cushing et al.
2005; Rayner et al. 2009) of brown dwarfs with data
overlapping with our narrowband mid-IR filters ([3.09]
and [4.05]) spanning spectral classes between L1 and
L5: 2MASS J14392836+1929149 (L1), Kelu-1AB (L2),
2MASS J15065441+1321060 (L3), 2MASS J15074769-
1627386 (L5).
To compare the β Pic b pho-
tometry with cooler L dwarfs, we add combined
IRTF/SpeX and Subaru/IRCS spectra from 1 to 4.1
µm for 2MASS J08251968+2115521 (L7.5) and DENIS-
P J025503.3-470049 (L8) (Cushing et al. 2008). Finally,
we add spectra for the low surface-gravity L4.5 dwarf,
2MASSJ22244381-0158521 (Cushing et al. 2008). To
highlight differences between β Pic b and these L dwarfs,
we scale the flux densities for each of these standards to
match β Pic b at ∼ 2.15 µm (Ks band).
To convert our photometry derived in magnitudes to
flux density units, we use the zeropoint fluxes listed in
Table 6. The JHKs and L′M ′(4.78 µm) zeropoints
are from Cohen et al. (2003) and Tokunaga and Vacca
(2005), respectively. We base the other zeropoints off
of Rieke et al. (2008), although alternate sources (e.g.
Cohen et al. 1995) yield nearly identical values. Because
the overlap in wavelengths between β Pic and these ob-
jects is not uniform, we do not perform a rigorous fit
between the two, finding the scaling factor that mini-
mizes the χ2 value defined from the planet flux density,
comparison object flux density, and photometric errors
in both. Rather, we focus on a simple first-order com-
parison between β Pic b and the comparison objects to
motivate detailed atmospheric modeling later in Section
4.
Figure 7 (left panel) compares photometry for β Pic b
to spectra for field L1 -- L5 dwarfs. While the L1 standard
slightly overpredicts the flux density at J band, the other
three early/mid L standards match the β Pic b near-
IR SED quite well, indicating a "near-IR spectral type"
of ∼ L2 -- L5. The L7.5 and L8 standards also produce
reasonable matches, although they tend to underpredict
the brightness at J band (right panel).
However, all standards have difficulty matching the β
Pic b SED from 3 -- 4 µm. In particular, the β Pic b flux
density from ∼ 3 to ∼ 5 µm is nearly constant, whereas it
rises through 4 µm and then steeply drops in all six stan-
dards depicted here. Focused on only β Pic b photom-
etry at 3.8 -- 4.1 µm, the "mid-IR spectral type" is hard
6
to define, the low surface gravity L4.5 dwarf bears the
greatest resemblance, although we fail to identify good
matches at all wavelengths with any of our spectral tem-
plates, where the 3.1 µm, L′, and [4.05] data points are
the most problematic. While none of our standards have
measurements fully overlapping with the M ′ filter, the
flux densities at 5.1 µm indicate that they may have a
very hard time simultaneously reproducing our measure-
ments at all four filters between 3 and 5 µm. Although
non-equilibrium carbon chemistry can flatten the spectra
of low surface gravity L/T dwarfs (Skemer et al. 2012),
its effect is to weaken the methane absorption trough at
∼ 3.3 µm and suppress emission at ∼ 5 µm. Thus, it
is unclear whether this effect can explain the enhanced
emission at ∼ 3.1 µm (mostly outside of the CH4 ab-
sorption feature to begin with) and 5 µm.
To understand whether β Pic b's SED is unique even
amongst other very low-mass substellar objects, we com-
pare our photometry to that for companions listed in
Table 4 that have photometry from 1 µm through ∼ 4 -- 5
µm: HR 8799 bcd, κ And b, 1RXJ 1609 B, GSC 06214B,
HIP 78530 B, 2M 1207A/B, HR 7329B, and AB Pic. Two
objects -- 1RXJ 1609 B and GSC 06214B -- have 3.1 µm
photometry: 1RXJ 1609 B from (Bailey et al. 2013) and
κ And b has [4.05] from data obtained by T. C. (M[4.05]
= 9.45 ± 0.20) (Bonnefoy, Currie et al., 2013 in prep.).
The two far-right columns of Table 5 lists the re-
duced χ2 and goodness-of-fit statistics between β Pic
b's JHKsL′ ([3.1],[4.05]) photometry, while Figure 8
displays these comparisons for κ And b, 1RXJ 1609B,
and GSC 06214B, which are all thought to be low sur-
face gravity companions with Tef f ∼ 1700 K, 1800 K,
and 2200 K (Carson et al. 2013; Lafreni`ere et al. 2010;
Bowler et al. 2011; Bailey et al. 2013). Overall, κ And
b provides the best match to β Pic b's photometry, re-
quires negligible flux scaling, and is essentially the same
within the 68% confidence limit (C.L.) (χ2 = 0.946, C.L.
= 0.186), although the large photometric uncertainties
in the near-IR limit the robustness of these conclusions.
The companion to 1RXJ 1609 likewise produces a very
good match (χ2 = 1.369, C.L. = 0.287), while the slightly
more luminous (and massive) GSC 06214B appears to be
much bluer, (relatively) overluminous in L′ and M ′ (or,
conversely, overluminous at JHKs) by ∼ 30%. In com-
parison, the cooler (Tef f ≈ 900-1100 K) exoplanets HR
8799 bcd provide far poorer matches (χ2 ∼ 6 -- 52).
Still, it is unclear whether any object matches β Pic b's
photometry at all wavelengths: both of the objects for
which we have [3.1] data, GSC 06214B and 1RXJ 1609B,
are still slightly underluminous here. Moreover, the best-
matching companions -- κ And b and 1RXJ 1609B -- are
still not identical, as the scaling factors between β Pic b's
spectrum and these companions' spectra that minimize
χ2 are ∼ 0.83 and 0.53, respectively. While compan-
ions with identical temperatures but radii 10% and 30%
larger than β Pic b would achieve this scaling, κ And
b and 1RXJ 1609B are respectively older and younger
than β Pic b, whereas for a given initial entropy of for-
mation planet radii are expected to decrease with time
(Spiegel and Burrows 2012).
In summary, young (low surface gravity?), low-mass
objects may provide a better match to β Pic b's pho-
tometry than do field dwarfs, especially those with tem-
peratures well above 1000 K but slightly below 2000 K
(κ And b, 1RXJ 1609 B). However, we fail to find a
match (within error bars) between the planet's photom-
etry spanning the full range of wavelengths for which we
have data, especially at ∼ 3 µm. As the planet spectra
depend critically on temperature, surface gravity, clouds
and (as we shall see) dust particle sizes, our comparisons
imply that β Pic b may differ from most young substellar
objects in one of these respects. Next, we turn to detailed
atmospheric modeling to identify the set of atmospheric
parameters that best fit the β Pic b data.
4. PLANET ATMOSPHERE MODELING
To further explore the physical properties of β Pic b,
we compare its photometry to planet atmosphere models
adopting a range of surface gravities, effective tempera-
tures, and cloud prescriptions/dust. For a given sur-
face gravity and effective temperature, a planet's emitted
spectrum depends primarily on the atmosphere's com-
position, the structure of its clouds, and the sizes of
the dust particles of which the clouds are comprised
(Burrows et al. 2006). For simplicity, we assume solar
abundances except where noted and leave consideration
of anomalous abundances for future work.
Based on β Pic b's expected luminosity (log(Lp/L⊙)
∼ -3.7 to -4, Lagrange et al. 2010; Bonnefoy et al. 2013)
and age, it is likely too hot (Tef f ∼ 1400-1800 K) for
non-equilibrium carbon chemistry to play a dominant
role (Hubeny and Burrows 2007; Galicher et al. 2011).
Therefore, our atmosphere models primarily differ in
their treatment of clouds and the dust particles entrained
in clouds. For each model, we explore a range of surface
gravities and effective temperatures.
4.1. Limiting Cases: The Burrows et al. (2006) E60
and A60 Models and AMES-DUSTY Models
4.1.1. Model Descriptions
We begin by applying an illustrative collection of
previously-developed atmosphere models to β Pic b.
These models will produce limiting cases for the planet's
cloud structure and typical dust grain size, which we re-
fine in Section 4.2. To probe the impact of cloud thick-
ness, we first adopt a (large) modal particle size of 60 µm
and consider three different cloud models: the standard
chemical equilibrium atmosphere thin-cloud models from
Burrows et al. (2006), which successfully reproduces the
spectra of field L dwarfs, moderately-thick cloud models
from Madhusudhan et al. (2011), and thick cloud mod-
els used in Currie et al. (2011a). To investigate the im-
pact of particle size, we then apply the AMES-DUSTY
models. The DUSTY models lack any dust grain sed-
imentation, such that the dust grains are everywhere
in the atmosphere, similar to the distribution of dust
grains entrained in thick clouds. However, they adopt
far smaller dust grains than do the thick cloud mod-
els from Madhusudhan et al. (2011) and Currie et al.
(2011a), where the grains are submicron in size and fol-
low the interstellar grain size distribution (Allard et al.
2001). All models described here and elsewhere in the
paper assume that the planet is in hydrostatic and ra-
diative equilibrium. None of them consider irradiation
from the star, as this is likely unimportant at β Pic b's
orbital separation. Table 7 summarizes the range of at-
mospheric properties we consider for each model.
The Burrows et al. (2006) E60 Thin Cloud,
Large Dust Particle Models -- As described in
Burrows et al. (2006) and later works (e.g. Currie et al.
2011a; Madhusudhan et al. 2011), the Model E60 case
assumes that the clouds are confined to a thin layer,
where the thickness of the flat part of the cloud encom-
passes the condensation points of different species with
different temperature-pressure point intercepts. Above
and below this flat portion, the cloud shape function de-
cays as the -6 and -10 powers respectively, so that the
clouds have scale heights of ∼ 1/7th and 1/11th that of
the gas. We adopt a modal particle size of 60 µm and
a particle size distribution drawn from terrestrial water
clouds (Deirmendjian 1964). We consider surface grav-
ities with log(g) = 4 and 4.5 and temperatures with a
range of Tef f = 1400 -- 1800 K in increments of 100 K.
(2011)
The Madhusudhan et al.
AE60
Moderately-Thick Cloud, Large Dust Parti-
cle Models -- Described in Madhusudhan et al. (2011),
the Model AE60 case assumes a shallower cloud shape
function of su = 1, such that the cloud scale height
is half that of the gas as a whole. We again adopt a
modal particle size of 60 µm and the same particle size
distribution. We consider surface gravities with log(g) =
4 and 4.5 and temperatures between Tef f = 1000 -- 1700
K in increments of 100 K.
The Burrows et al. (2006) A60 Thick Cloud,
Large Dust Particle Models -- As described in
Currie et al. (2011a), the Model A60 case differs in that
it assumes that the clouds extend with a scale height
that tracks that of the gas as a whole. Below the flat
part of the cloud, the shape function decays as the -10
power as in the E60 and AE60 models, although devi-
ations from this do not affect the emergent spectrum.
Here, we consider surface gravities with log(g) = 4 and
4.5 and temperatures with a range of Tef f = 1000-1700
K in increments of 100 K.
AMES-DUSTY Thick-Cloud, Small Dust Par-
ticle Limit -- The AMES-DUSTY atmosphere models
(Allard et al. 2001) leverage on the PHOENIX radiative
transfer code (Hauschildt and Baron 1999) and explore
the limiting case where dust grains do not sediment/rain
out in the atmosphere.
Unlike the Burrows et al.
(2006) models and those considered in later works (e.g.
Spiegel and Burrows 2012), the AMES-DUSTY models
adopt a interstellar grain size distribution favoring far
tinier dust grains with higher opacities. The grains'
higher opacities reduce the planet's radiation at shorter
wavelengths. Thus, these models have dramatically
different near-IR planet spectra from the E/A/AE60
type models with larger modal grain sizes even at the
same temperatures and gravities (cf. Burrows et al. 2006;
Currie et al. 2011a). Here we consider AMES-DUSTY
models with log(g) = 3.5, 4, and 4.5 and Tef f = 1000 --
2000 K (∆Tef f =100 K).
4.1.2. Fitting Method
To transform the DUSTY spectra into predicted flux
density measurements (at 10 pc), we convolve the spec-
tra over the filter response functions and scale by a di-
lution factor of f = (Rplanet/10 pc)2. We consider a
range of planet radii between 0.9 RJ and 2 RJ . Like-
wise, we convolve the E60 and A60 model spectra over
filter response functions. The E60 models (as do all
7
other Burrows et al. 2006 and Madhusudhan et al. 2011
models) adopt a mapping between planet radius and
surface gravity/temperature set by the Burrows et al.
(1997) planet evolution models. To explore departures
from these models, we allow the the radius to vary by an
additional scale factor of 0.7 to 1.7. For most of our grid,
this translates into a radius range of 0.9 to 2 RJ .
Our atmosphere model fitting follows methods in
Currie et al. (2011a,b), where we quantify the model fits
with the χ2 statistic,
χ2 =
n
X
i=0
(fdata,i − Fmodel,i)2/σ2
data,i.
(1)
We weight each datapoint equally. Because our pho-
tometric calibration fully considers uncertainties due to
the signal-to-noise ratio, the processing-induced attentu-
ation, and the absolute photometric calibration, we do
not set a 0.1 mag floor to σ for each data point as we
have done previously.
We determine which models are formally consistent
with the data by comparing the resulting χ2 value to that
identifying the 68% and identify those that can clearly
be ruled out by computed the 95% confidence limit.
Note here that these limits are significantly more strin-
gent compared to the ones we adopted in Currie et al.
(2011a). Treating the planet radius as a free parameter,
we have five degrees of freedom for seven data points,
leading to χ2
68% = 5.87 and χ2
95% = 11.06. '
4.1.3. Results
Table 8 summarizes our fitting results using the E60,
AE60, A60, and DUSTY models. Figure 9 displays some
of these fitting results, where the left-hand panels show
the χ2 distributions with the 68% and 95% confidence
limits indicated by horizontal lines dashed and dotted
lines. The right-hand panels and middle-left panel show
the best-fitting models for each atmosphere prescription.
A successful model must match three key properties of
the observed SED: (1) At 3 -- 5 µm, the SED is relatively
flat, (2) at 1 -- 3 µm, the spectral slope is relatively shal-
low, and (3) the overall normalization of the 3 -- 5 µm flux
relative to the 1 -- 3 µm flux must match the data.
For the E60, AE60, and A60 models, we find χ2 min-
ima at log(g) = 4 -- 4.5 and Tef f = 1400 K in each case
with radius scaling factors, the constant we multiple the
nominal Burrows et al. planet radii, between 1.185 and
1.680. For the Burrows et al. (1997) evolutionary mod-
els, these scaling factors imply planet radii between ∼
1.8 and 2 RJ , at the upper extrema of our grid in radius.
Figure 9 illustrates the impact on the SED of chang-
ing cloud models, given a fixed grain size. The best-fit
temperature does not vary dramatically because, roughly
speaking, the relative fluxes at 1 -- 3 µm and 3 -- 5µm are
determined by the SED's blackbody envelope. However,
cloud thickness dramatically affects the depths of absorp-
tion bands superimposed on that envelope. Atmosphere
models presented here do not feature temperature inver-
sions. As such, high opacity molecular lines have low
flux densities because they originate at high altitudes
where the temperature is low. When clouds are thin, op-
tical depth unity is achieved at very different altitudes in
and outside of absorption bands such as those at 3.3µm
8
(methane) and 4.5 µm (primarily CO), and the bands
appear deep.
For a fixed observed effective temperature, thicker
clouds translate into hotter temperature profiles (i.e. at
a given pressure in the atmosphere, the temperature is
higher) (e.g. Madhusudhan et al. 2011). The total Rosse-
land mean optical depth of the atmosphere at a given
pressure is higher (Madhusudhan et al. 2011). As the
clouds become thicker, the τ = 1 surface also is more uni-
form, such that molecular features wash out and the spec-
trum overall appears flatter and more like a blackbody
(Burrows et al. 2006). Hence, the prominent molecular
absorption bands seen in the best-fit E60 (thin cloud)
model are substantially reduced in the A60 (thick cloud)
model, with AE60 lying in between. The planet's flat
3 -- 5 µm SED is best fit by A60.
Although the χ2 minima for all four of the models we
consider are sharply peaked, none yield fits falling within
the 68% confidence interval. The fits from E60 and AE60
are particularly poor, ruled out at a greater than 5-σ
level, whereas the A60 model quantitatively does bet-
ter but still is ruled out as an acceptably-fitting model
(C.L. ∼ 3.9-σ). The best-fit AMES-DUSTY model fits
the SED even better than A60, with parameters of Tef f
= 1700 and log(g) = 3.5 and a radius of r = 1.35 RJ , sim-
ilar parameters to those found in Bonnefoy et al. (2013).
However, the best-fit DUSTY model still falls outside
the 68% confidence limit (C.L. = 0.84). These exercises
suggest that the atmospheric parameters assumed in the
models need to be modified in order to better reproduce
the β Pic b photometry. To achieve this, we restrict
ourselves to thick clouds and consider more carefully the
impact of dust size.
4.2. A4, Thick Cloud/Small Dust Models
4.2.1. The Effect of Small Dust Particles
Our analyses in the previous section show the extreme
mismatch between standard L dwarf atmosphere models
assuming thin clouds and large dust particles and the
data. While our χ2 values for the Burrows thick cloud,
large dust particle models are systematically much lower,
they likewise are a poor match to the data. In contrast,
fits from the AMES-DUSTY models only narrowly lie
outside the 68% confidence interval.
A closer inspection of the best-fitting models in each
case (right-hand panels) illustrates how they fail. The
main difficulty with matching these models to β Pic b
spectrum is the planet's flat SED from 2 µm to 5 µm,
where models tend to underpredict the flux density at
3.1 µm and/or M ′. The slope from J to Ks is also a
challenge. Reducing dust sizes can further fill in absorp-
tion troughs by increasing the opacities of the clouds.
The AMES-DUSTY model, however, appears to over-
correct as its spectrum exhibits sharp peaks due to its
submicron sized grains that degrade its fit to the data.
Therefore, we consider grain sizes intermediate between
those in A60 and AMES-DUSTY (e.g. ∼ 1 -- 30 µm).
A4 Thick Cloud, Small Dust Particle Models --
As the primary difference between these models is the
typical/modal particle size, we here introduce a new set
of atmosphere models with the same A-type, thick cloud
assumption but with modal particle sizes slightly larger
than those characteristic of dust in the AMES-DUSTY
models but significantly smaller than previous Burrows
models. We nominally adopt 4 µm as our new modal
particle size, comparable in wavelength to the peak flux
density of β Pic b in Fν units. Thus, we denote these
models as "A4", thick-cloud, small dust particle models.
Figure 10 illustrates the effect of dust on the planet
spectrum for modal particle sizes of 3, 5, 30 and 50 µm
and a temperature and surface gravity consistent with
that expected to reflect β Pic b based on planet cool-
ing models (Tef f = 1600 K, log(g)=3.8-4, r ∼ 1.5 RJ )
(Burrows et al. 1997; Baraffe et al. 2003; Lagrange et al.
2010; Spiegel and Burrows 2012; Bonnefoy et al. 2013).
As particle sizes decrease, the water absorption troughs
at 1.8 µm and 2.5 µm diminish. Likewise filled in
is the deep absorption trough at ∼ 3.3 µm and 4.5
µm that is usually diagnostic of carbon chemistry (e.g.
Hubeny and Burrows 2007; Galicher et al. 2011). Over-
all, the spectrum flattens and becomes redder (shorter
wavelength emission originates at higher altitudes), with
weaker emission and a steeper slope at J to Ks. This
reddening explains the difference in best-fit effective tem-
perature between the AMES-DUSTY model and the 60
µm dust models.
4.2.2. Model Fitting Procedure
We follow the steps outlined in Currie et al. (2011a),
where we perform two runs: one fixing the planet radius
to the Burrows et al. (1997) hot-start predictions for a
given Tef f and log(g) and another where we consider a
range of planet radii (as in the previous section). For the
fixed-radii modeling, the 68% and 95% confidence limits
now lie at χ2 = 7.01 and 12.6, respectively, whereas they
are at 5.87 and 11.06 for the varying-radii fits as before.
Similar to the Burrows A/E60 model runs, we consider
a range of temperatures between 1400 K and 1900 K. To
explore whether or not the fits are sensitive to surface
gravity, we consider models with log(g) = 3.6, 3.8, 4,
and log(g) = 4.25. For the age of β Pic (formally, 8 to 20
Myr), this surface gravity range fully explores the masses
(in the hot-start formalism) allowed given the radial-
velocity dynamical mass limits (Lagrange et al. 2012a).
To further explore the effect that carbon chemistry
may have on our planet spectra, we take the best-fitting
model from the above exercise, significantly enhance the
methane abundances over solar and re-run a small grid
of temperatures based on that, to determine if depar-
tures from solar abundances may yield a wider range of
acceptable atmosphere parameters. Because variations
in molecular abundances affect the depths of molecular
absorption bands, we expect that such variations may
improve our fit.
4.2.3. Results
Figures 11 and 12 and Table 8 present our results for
fitting the β Pic b data with the A4, thick cloud/small
dust models. Quantitatively, these models better repro-
duce the β Pic b SED. Fixing the planet radius to val-
ues assumed in the Burrows et al. (1997) planet cooling
model, we find one atmosphere model -- log(g) = 3.8, Tef f
= 1600 K -- consistent with the data to within the 68%
confidence interval. A wide range of models are consis-
tent with the data at the 95% confidence limit, covering
± 0.2 dex in surface gravity and ± 100 K in temperature.
We can slightly improve upon these fits if we allow
the planet radius to freely vary. In this case, the best-
fitting models yield a slightly higher surface gravity of
log(g) = 4 -- 4.25 but the same temperature of 1600 K.
But in contrast to the fixed-radius case above, a wide
range of models are consistent with the data at the 68%
confidence limit. In particular, all surface gravities con-
sidered in our model grid are consistent with the data
provided that the temperature is 1600 K and the radius
is rescaled accordingly: log(g) = 3.6 -- 4.25, Tef f = 1600
K. Another set of models with the full range of surface
gravities and 250 K spread in temperature (1500 -- 1750
K) are marginally consistent with the data.
The methane-enhanced models are shown in Figure 13
for log(g)=4 and Tef f = 1575 -- 1650 K. The 1575 K and
1600 K models (Figure 13) likewise produce good fits
to the data (χ2 = 5.13 -- 5.3), where the 1650 K model
barely misses the 68% cutoff. Thus, while best-fitting
solar abundance models appear narrowly peaked at Tef f
= 1600 K, the range in temperature enclosing the 68%
confidence interval is larger when non-solar abundances
are considered, at least extending from 1575 K to almost
1650 K. Changes in molecular abundances, as expected,
allow us to very slightly improve the SED fit. However,
thick clouds and small dust grains are likely still needed
to match the emission from β Pic b, since given molecules
(i.e. CH4) by themselves do not change fluxes compara-
bly at 1 -- 3 µm and 3 -- 5 µm.
In summary, adopting the Burrows et al. (1997) hot-
start models to set our planet radii and the A4 thick
cloud/small dust atmosphere models, we derive log(g) =
3.8 and Tef f = 1600 K for β Pic b. Allowing the radius
to vary and considering non-solar carbon abundances we
derive log(g) = 3.6 -- 4.25 and Tef f = 1575 -- 1650 K, mean-
ing that the planet temperature is well constrained but
the surface gravity is not. However, in Section 5 we nar-
row the range of surface gravities to log(g) = 3.8 ± 0.2,
as higher surface gravities imply planet masses ruled out
by dynamical estimates.
4.2.4. Varying Grain Sizes and Fits Over Other Model
Parameter Space
The models considered in the previous subsections as-
sume thick clouds, dust grains with a modal size of 4
µm, and (in most cases) solar abundances. Although we
achieve statistically significant fits to the β Pic b pho-
tometry with these models, our exploration of model pa-
rameter space is still limited. While an exhaustive pa-
rameter space search is beyond the scope of this paper,
here we argue that models either thick clouds or small
dust grains are unlikely to produce good-fitting models.
Thus, small grains and thick clouds are likely important
components of β Pic b's atmosphere required in order to
fit the planet's spectrum.
To consider the robustness of our results concerning
the modal grain size, we also ran some model fits for
modal particle sizes of 3 µm, 5 µm, 10 µm, and 30 µm.
The models with 3 and 5 µm modal sizes yielded fits
slightly worse than those with modal sizes of 4 µm. For
example, models with modal sizes of < a > = 3 and 5
µm, Tef f = 1600 K, log(g) = 3.8 and a freely-varying
planet radius yield χ2 6.31 and 6.28, respectively. These
values lie slightly outside the 68% confidence interval,
although they are still smaller than those from the best-
9
fit DUSTY models. In contrast, models with < a > =
10 µm and 30 µm fit the data significantly worse (χ2 =
10.0 and 19.6, respectively).
Similarly, our investigations show that small dust
grains do not obviate the need to assume thick, A-type
clouds in our atmosphere models. For example, adopting
the AE-type cloud prescription, modal particle sizes of
5 µm, a temperature of Tef f = 1600 K, and a surface
gravity of log(g) = 3.8 -- 4, our model fits are substanti-
ailly worse than the A4-type models and even the AMES-
DUSTY models and are easily ruled out (χ2 ∼ 15 -- 40).
The AE-type cloud prescription fails to reproduce the β
Pic b spectrum because by confining clouds to a thinner
layer the τ = 1 surface varies too much in and out of
molecular absorption features such as CH4 and CO. In
disagreement with the β Pic b SED, the AE model spec-
tra thus have suppressed emission at ≈ 3 µm and 5 µm
and an overall shape looking less like a blackbody.
In contrast, non-solar abundances may slightly widen
the range of parameter space (in radius, temperature,
gravity, etc.) yielding good fits. The methane-rich model
from the previous section adopting < a > = 5 µm instead
of 4 µm, log(g) = 3.8, and Tef f = 1600 K still yields a fit
in agreement with the data to within the 68% confidence
limit (χ2 = 5.59). Thus, within our atmosphere model-
ing approach we need 1) grains several microns in size,
comparable to the typical sizes of grains in debris disks,
and 2) thick clouds to yield fits consistent with the data
to within the 68% confidence limit. These results are
not strongly sensitive to chemical abundances although
varying the range of abundances may slightly widen the
corresponding range of other parameter space (in tem-
perature, gravity, etc.) yielding good-fitting models.
5. PLANET RADII, LUMINOSITIES, MASSES, AND
EVOLUTION
From the set of models that reproduce the β Pic b
SED to the 68% confidence interval, we derive a range
of planet radii, luminosities and inferred masses. The
planet radii for each model run are given in Table 8. In-
terestingly, all of our 1-σ solutions fall on or about R
∼ 1.65 RJ with very little dispersion (± ∼ 0.05 dex).
If we consider the range of radii for a given atmosphere
model consistent with the data to within the 68% (or
95%) confidence interval regardless of whether the given
radius is the best-fit one, then the range in accept-
able radii marginally broadens: r = 1.65 ± 0.06. Note
that these radii are larger than those inferred for HR
8799 bcde based either on its luminosity and hot-start
cooling tracks (Marois et al. 2008, 2010a) or from at-
mosphere modeling, where in Currie et al. (2011a) and
Madhusudhan et al. (2011) our best-fit models typically
had R ∼ 1.3 RJ . The range in inferred planet luminosi-
ties is even narrower, The values inferred from our best-
fit models center on log(L/L⊙)= -3.80 with negligible
intrinsic dispersion (± 0.01 dex). The uncertainty in β
Pic's distance affects both our radius and luminosity de-
terminations. Treating the distance uncertainty (± 1 pc)
as a separate, additive source of error, β Pic b's range in
radii is 1.65 ± 0.06 RJ and its luminosity is log(L/L⊙)=
-3.80 ± 0.02.
From our best-fit surface gravities and inferred radii,
we can derive the mass of the planets inferred from
our modeling. Adopting the hot-start formalism with-
10
out rescaling the radius, our modeling implies a best-
fit planet mass of ∼ 7 MJ ; the range covering the 95%
confidence limit of 5 -- 9 MJ .
If we allow the radius to
freely vary, we derive a range of 4 MJ to 18.7 MJ ,
where the spread in mass reflects primarily the spread
in surface gravity from best-fitting models (log(g) = 3.6 --
4.25). However, RV data limits β Pic b's mass to be
less than 15 MJ if its semimajor axis is less than 10
AU , which appears to be the case (Lagrange et al. 2012a;
Chauvin et al. 2012; Bonnefoy et al. 2013). Thus, limit-
ing the atmosphere models to those whose implied masses
do not in violate the RV upper limits (ones with log(g) =
3.6 -- 4), our best-estimated (68% confidence limit) planet
masses are ∼ 7+4
−3 MJ.
Planets cool and contract as a function of time, and
we can compare our inferred luminosities and radii to
planet cooling models. Figure 14 compares the inferred
planet luminosity to the hot-start planet evolution mod-
els from Baraffe et al. (2003). For context, we also show
the luminosities of other 5 -- 100 Myr old companions with
masses that (may) lie below 15 MJ : GSC 06214 B,
1RXJ 1609 B, HR 8799 bcde, AB Pic B, and κ And b.
From our revised luminosity estimate, the Baraffe et al.
(2003) hot-start models imply a mass range of ∼ 8 -- 12
MJ if the planet's age is the same as the star's inferred
age (12+8
If we use the
Burrows et al. (1997) hot-start models, we obtain nearly
identical results of 9 -- 13 MJ . These masses are slightly
higher than most of the implied masses from our atmo-
sphere modeling but still broadly consistent with them
and with the dynamical mass upper limits of 15 MJ from
Lagrange et al. (2012a). Note also that the luminosities
and planet radii are completely inconsistent with pre-
dictions from low-entropy, cold-start models for planet
evolution.
−4 Myr; Zuckerman et al. 2001).
Still, the righthand panel of Figure 14 highlights one
possible complication with our results, namely that our
best-estimated planet radii are near the upper end of
the predicted range for 5 -- 10 MJ companions in the hot-
start formalism. For the hot-start models presented
in Burrows et al. (1997) and Baraffe et al. (2003), 5 --
10 MJ companions are predicted to have radii of ∼
1.5 -- 1.6 RJ .
For the hot-start models presented in
Spiegel and Burrows (2012), the predicted range for 5 -- 10
MJ planets covers ≈ 1.4 -- 1.5 MJ
13.
To reduce the planet radius of ∼ 1.65 RJ by ∼ 10%
while yielding the same luminosity requires raising the
effective temperature from ≈ 1600 K to ∼ 1700 K. This
is a small change and atmospheric modeling of β Pic b
and similar substellar objects is still in its early stages.
Thus, it is quite plausible that future modeling efforts,
leveraging on additional observations of β Pic b and those
of other planets with comparable ages and luminosities,
will find quantitatively better fitting solutions that im-
ply smaller planet radii and higher temperatures. We
consider this to be the most likely explanation.
13 This mismatch does not mean that the AMES-DUSTY mod-
els, whose fits to the data imply planet radii of ≈ 1.3 RJ and lie
just outside the 68% confidence limit, are preferable. The best-
fit AMES-DUSTY radii lie below the radii predicted for 5 -- 10 MJ
objects at β Pic b's age and are only consistent for 'warm-start'
models that imply lower luminosities and colder temperatures than
otherwise inferred from the AMES-DUSTY fits.
Alternatively, we can bring the atmosphere modeling-
inferred radius into more comfortable agreement with
hot-start evolutionary models if β Pic b is ≈ 7 Myr old
or less. For a system age of ≈ 12 Myr, this is consistent
with it forming late in the evolution of the protoplan-
etary disk that initially surrounded the primary. Even
adopting the lower limit on β Pic's age (8 Myr), β Pic
b may still need to be younger than the star. While
most signatures of protoplanetary disks around 1 -- 2 M⊙
stars disappear within 3 -- 5 Myr, some ∼ 10 -- 20% of such
stars retain their disks through 5 Myr (Currie et al. 2009;
Currie and Sicilia-Aguilar 2011; Fedele et al. 2010). Sev-
eral 1 -- 2 M⊙ members of Sco-Cen and h and χ Persei ap-
parently have even retained their disks for more than 10
Myr (Pecaut et al. 2012; Bitner et al. 2010; Currie et al.
2007), comparable to or greater than the age of β Pic.
Models for even rapid planet formation by core accre-
tion predict that several Myrs elapse before the cores
are massive enough to undergo runaway gas accretion
at β Pic-like separations (Kenyon and Bromley 2009;
Bromley and Kenyon 2011).
In Figure 14 the open circles depict a case where β
Pic b formed after 5 Myr, effectively making the planet
5 Myr younger than the star, where the implied masses
and radii overlap better with our atmospheric modeling-
inferred values. The overlap is even better for some hot-
start models such as COND, which predict larger planet
radii at ≈ 5 -- 10 Myr than depicted here. Note that a
young β Pic b as depicted in Figure 14 with an implied
mass mass of M ≥ 5 MJ is still consistent with a scenario
where the planet produces the warped secondary disk
(c.f. Dawson et al. 2011).
6. DISCUSSION
6.1. Summary of Results
This paper presents and analyzes new/archival
VLT/NaCo and Gemini/NICI 1 -- 5 µm photometry for
β Pictoris b, These data allow a detailed comparison be-
tween β Pic b's SED and that of field brown dwarfs and
other low-mass substellar objects such as directly imaged
planets/candidates around HR 8799 and κ And. Using
a range of planet atmosphere models, we then constrain
β Pic b's temperature, surface gravity and cloud proper-
ties. Our study yields the following primary results.
• 1. - The near-IR (JHKs) colors of β Pic b appear
fairly consistent with the field L/T dwarf sequence.
Compared to other young, low-mass substellar ob-
jects, β Pic b's near-IR colors bear the most resem-
blance to late M to early T dwarfs such as Luh-
man 16B and κ And b. From its near-IR colors
and color-magnitude positions, β Pic b's near-IR
properties most directly resemble those of a L2 -- L5
dwarf.
• 2. - β Pic b's mid-IR properties identify a signifi-
cant departure from the field L/T dwarf sequence.
The planet is slightly overluminous at L′ and signif-
icantly overluminous at M ′, with deviations from
the field L dwarf sequence matched only by GSC
06214B and κ And b. The mid-IR portion of β Pic
b's SED appears more like that of a late L dwarf
or low surface gravity mid L dwarf. The broad-
band JHKsL′ photometry for β Pic b also closely
resembles that of κ And b. However, it is unclear
whether any object matches β Pic b's SED at all
wavelengths for which we have measurements. Its
3.1 µm brightness and 3.8 -- 5 µm spectral shape are
particularly difficult to match.
• 3. -- Compared to limiting-case atmosphere mod-
els E60 (large dust confined to very thin clouds),
AE60/A60 (large dust confined to moderately-
thick/thick clouds) and DUSTY (copious small
dust everywhere in the atmosphere), β Pic b ap-
pears to have evidence for thick clouds consistent
with a high Tef f and low surface gravity. We fail to
find any E60/AE60/A60 model providing statisti-
cally significant fits over a surface gravity range of
log(g) = 4 -- 4.5 and any Tef f . The DUSTY models
come much closer to yielding statistically signifi-
cant fits but mismatch the planet flux at J, Ks,
[3.1], and M ′. From these fiducial comparisons, we
infer that β Pic b's atmosphere shows evidence for
clouds much thicker than those assumed in the E60
models but is slightly less dusty than the DUSTY
models imply.
• 4. -- Using thick cloud models with particle sizes
slightly larger than those found in the interstellar
medium (< a > = 4 µm), we can match β Pic b's
SED in both the near and mid IR. Assuming planet
radii appropriate for the Burrows et al. (1997) 'hot-
start' models, we derive log(g) = 3.80 and Tef f =
1600 K for β Pic b. Allowing the radius to freely
vary, leaves the surface gravity essentially uncon-
strained, where models consistent with the data at
the 68% confidence limit include log(g) = 3.6 -- 4.25
and Tef f = 1600 K. Considering departures from
solar abundances and eliminating models that im-
ply masses ruled out by dynamical estimates, the
acceptably fitting range of atmosphere parameters
cover log(g) = 3.6 -- 4 and Tef f = 1575 -- 1650 K.
• 5. -- Using our best-fit atmosphere models and
eliminating models inconsistent with β Pic b's dy-
namical mass upper limit, within the hot-start for-
malism we derive a mass of 7 MJ for a fixed ra-
dius and 7+4
−3 MJ for a scaled radius. Our best-fit
planet radius is ∼ 1.65 ± 0.06 RJ and luminosity
of log(L/L⊙) = -3.80 ± 0.02.
• 6. -- While our derived luminosity and radius for
β Pic b rules out cold start models, the radius is
near the upper end of predicted radii for hot start-
formed planets with β Pic's age. As the planet only
needs to be ∼ 100 K hotter to easily eliminate this
discrepancy, it likely identifies a limitation of the
atmosphere models. Alternatively, if β Pic b has a
significantly younger age than the star's age con-
sistent with it forming late in the protoplanetary
disk stage our derived radius is comfortably within
the range predicted by hot start models.
6.2. Comparisons to Other Recent β Pictoris b Studies
6.2.1. Currie et al. 2011b
In our first-look analysis of the atmosphere of β Pic-
toris b (Currie et al. 2011b), we compared its Ks, L′,
11
and [4.05] photometry to an array of atmosphere mod-
els, from atmospheres completely lacking clouds to those
with the Model A-type thick clouds that extend to
the visible surface of the atmosphere.
In that pa-
per, we found that the AE thick cloud models from
Madhusudhan et al. (2011) yielded the smallest χ2 value.
The fits degraded at about the same level for the Model A
thick cloud and Model E "normal" L dwarf atmosphere
prescriptions, while the cloudless case fared the worst.
Currie et al. (2011b) conclude that while the AE thick
cloud model quantitatively produced the best fit, the ex-
isting data were too poor to say whether the clouds in β
Pictoris b were any different in physical extent, in mean
dust particle size, etc. from those for field L dwarfs with
the same range of temperatures.
Our present study greatly improves upon the analyses
in Currie et al. (2011b). First, our photometry covers
seven passbands, not three, at 1.25 -- 4.8 µm, not 2.18 -- 4.05
µm. This expanded coverage allows far firmer constraints
on β Pic b's atmospheric properties. In particular, our
new photometry strongly favors the Model A thick-cloud
prescription over AE, largely due to the relatively low
planet flux densities at 1.25 -- 1.65 µm and the relatively
high flux densities at 3.1 µm and M ′/4.8 µm, trends that
the Model A cases consistently reproduce better. While
all models considered in Currie et al. (2011b) assumed a
modal particle size of 60 µm for dust entrained in clouds,
our fits improve if we use smaller sizes. The combined
effect of thicker clouds and smaller particle sizes favor
atmosphere models with a slightly higher surface gravity
and temperature than the best-fit model in Currie et al.
(2011b). Our new data more clearly demonstrate the fail-
ure of the E models successful in fitting most of the field
L dwarf sequence and thus better distinguish β Pic b's
atmosphere from that of a typical cloudy field L dwarf.
6.2.2. Bonnefoy et al. (2013)
Bonnefoy et al. (2013) presented new photometry for
β Pictoris b in the J, H, and M ′ filters from data taken
in 2011 and 2012. The J and H detections are firsts
and greatly expand the wavelength coverage for β Pic
b's SED. Their M ′ detection is first well-calibrated detec-
tion, building upon and following the detection presented
in Currie et al. (2011b), which lacked a contemporane-
ous flux-calibration data to provide precise photometry.
They then combined these measurements with their pre-
viously published Ks and L′ photometry and [4.05] from
Quanz et al. (2010).
In general, our study clarifies and modifies, instead of
contradicting, the picture of β Pictoris b constructed in
Bonnefoy et al. (2013). On the same datasets, the SNR
of our detections is slightly higher but our photometry
agrees within theirs derived from their CADI, RADI, and
LOCI reductions within their adopted photometric un-
certainties (∼ 0.2 -- 0.3 mag). We derive smaller photo-
metric uncertainties, owing to a more uniform through-
put as a function of azimuthal angle, probably due to our
pixel masking technique and SVD cutoff in A-LOCI (see
also Marois et al. 2010b). We concur that the planet's
mid-IR colors are unusually red and highlight a poten-
tially strong, new disagreement with field L dwarfs at 3.1
µm.
We agree with Bonnefoy et al.'s general result that
the best-fitting atmosphere models are those interme-
12
diate between the AMES-DUSTY models (submicron-
sized dust everywhere) and the COND or BT-Settl mod-
els (no dust/clouds or very thin clouds). Quantitatively,
the χ2 values we derive are much larger than the best-
fitting models in Bonnefoy et al. because our photomet-
ric uncertainties are significantly smaller (e.g. 7 vs. 3 for
AMES-DUSTY). Our analyses point to thick clouds and
particle sizes small compared to the range typically used
in the Burrows et al. (2006) models but larger than the
ISM-like grains in the AMES-DUSTY models. The tem-
peratures, surface gravities, and luminosities they derive
are generally consistent with our best-fit values.
While they derive a lower limit to the initial entropy
of 9.3 kB/baryon, we do not provide a detailed simi-
lar analysis since the inferred entropy range depends on
the planet radius which, considering our studies together,
is very model dependent. Similarly, it depends on the
planet mass (for which there still is some range) and the
planet's age (which is very poorly constrained). Still, we
agree that cold start models are ruled out for β Pic b as
they fail to reproduce the inferred luminosity and radii
of the planet determined from both our studies.
6.3. Future Work to Constrain β Pic b's Properties
Deriving β Pic b's mass and other properties is diffi-
cult since they are based on highly uncertain parameters
such as the planet's age and its entropy at formation.
However, dynamical mass limits can be derived from
continued radial-velocity measurements (Lagrange et al.
2012a). As these limits depend on β Pic b's orbital pa-
rameters, future planet astrometry may be particularly
important in constraining β Pic b's mass.
If β Pic b
is responsible for the warp observed in the secondary
debris disk (Golimowski et al. 2006), planet-disk inter-
action modeling can likewise yield a dynamical mass es-
timate (Lagrange et al. 2009a; Dawson et al. 2011) pro-
vided the planet's orbit is known.
Finally, while our models nominally assume solar abun-
dances, we showed that changing the methane abundance
might yield marginally better fits to the data. Near-
infrared spectroscopic observations of β Pic b as can be
done soon with GP I and SP HERE may clarify its at-
mospheric chemistry. Future observations with GMT-
NIRS on the Giant Magellan Telescope should be capa-
ble of resolving molecular lines in β Pic b's atmosphere
(Jaffe et al. 2006), providing a more detailed look at its
chemistry, perhaps even constraining its carbon to oxy-
gen ratio and formation history (e.g Oberg et al. 2011;
Konopacky et al. 2013).
We thank Christian Thalmann, France Allard, and
the anonymous referee for helpful comments and discus-
sions and Michael Cushing for providing IRTF/SPeX and
Subaru/IRCS spectra of field L dwarfs. We are grate-
ful to the telescope staffs at ESO Paranal Observatory
and Gemini-South Cerro Pachon Observatory for sup-
port for our observations, all of which were obtained with
"delegated visitor mode" or "eavesdropping mode". Fi-
nally, we thank Christian Marois for very detailed dis-
cussions on image processing techniques and extensive
helpful suggestions that improved this manuscript. T.
C. acknowledges support from a McLean Postdoctoral
Fellowship. R. D. acknowledges NSF-GRFP grant DGE-
1144152.
Allard, F., et al., 2001, ApJ, 556, 357
Amara, A., Quanz, S., 2012, MNRAS, 427, 948
Baraffe, I., et al., 2003, A&A, 402, 701
Bailey, V., Hinz, P., Currie, T., et al., 2013, ApJ, 767, 31
Bejar, V. J., et al., 2008, ApJ, 673, L185
Beuzit, J.-L., et al., 2008, SPIE, 7014, 41
Biller, B., et al., 2010, ApJ, 720, 82L
Bitner, M., et al., 2010, 714, 1542
Boccaletti, A., et al., 2013, A&A, 551, L14
Bonnefoy, M., et al., 2010, A&A, 512, 52
Bonnefoy, M., et al., 2011, A&A, 528, 15L
Bonnefoy, M., et al., 2013a, A&A, 555, 107
Bowler, B., et al., 2010, ApJ, 733, 850
Bowler, B., et al., 2011, ApJ, 743, 148
Bromley, B., Kenyon, S. J., 2011, ApJ, 731, 101
Burgasser, A., et al., 2013, ApJ submitted, arXiv:1303.7283
Burrows, A., et al., 1997, ApJ, 491, 856
Burrows, A., et al., 2006, ApJ, 640, 1063
Carson, J., et al., 2013, ApJ, 763, L32
Chauvin, G., et al., 2004, A&A, 425, 29L
Chauvin, G., et al., 2005, A&A, 438, L29
Chauvin, G., et al., 2012, A&A, 542, 41
Cohen, M., et al., 1995, AJ, 110, 275
Cohen, M, et al., 2003, AJ, 126, 1090
Currie, T., et al., 2007, 669, L33
Currie, T., Lada, C. J, et al., 2009, ApJ, 698, 1
Currie, T., Bailey, V., et al., 2010, ApJ, 721, L177
Currie, T., Burrows, A., et al., 2011, ApJ, 729, 128
Currie, T., Thalmann, C., et al., 2011, ApJ, 736, L33
Currie, T., Sicilia-Aguilar, A., 2011, ApJ, 732, 24
Currie, T., et al., 2012a, ApJ, 760, L32
Currie, T., et al., 2012b, ApJ, 755, L34
Currie, T., et al., 2012c, ApJ, 757, 28
Cushing, M., et al., 2005, ApJ, 623, 1115
REFERENCES
Cushing, M., et al., 2008, ApJ, 678, 1372
Dawson, R., et al., 2011, ApJ, 743, L17
Deirmendjian, D., 1964, Appl. Opt., 3, 87
Delorme, P., et al., 2013, A&A, 553, L5
Fabrycky, D., Murray-Clay, R., 2010, ApJ, 710, 1408
Fedele, D., et al., 2010, A&A, 510, 72
Fortney, J., et al., 2008, ApJ, 683, 1104
Galicher, R., et al., 2011, ApJ, 739, L41
Galicher, R., Marois, C., et al., 2013, in prep.
Girard, J, et al., 2012, in Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference Series, Vol. 8447,
Society of Photo-Optical Instrumentation Engineers (SPIE)
Conference Series
Golimowski, D., et al., 2006, AJ, 131, 3109
Hauschildt, P., Baron, E., 1999, JCAM, 109, 41
Hubeny, I., Burrows, A., 2007, ApJ, 669, 1248
Ireland, M., et al., 2011, ApJ, 726, 113
Jaffe, D., et al., 2006, SPIE, 6269, 166
Kenyon, S., Bromley, B., 2009, ApJ, 690, L140
Konopacky, Q., et al., 2013, Science, 339, 1398
Kratter, K., et al., 2010, ApJ, 710, 1375
Kraus, A., Ireland, M., 2012, ApJ, 745, 1
Lafreni´ere, D., et al., 2007a, ApJ, 670, 1367
Lafreni´ere, D., et al., 2007b, ApJ, 660, 770
Lafreni´ere, D., et al., 2008, ApJ, 689, L153
Lafreni´ere, D., et al., 2010, ApJ, 719, 497
Lagrange, A.-M., et al., 2009, A&A, 493, L21
Lagrange, A.-M., et al., 2010, Science, 329, 57
Lagrange, A.-M., et al., 2012a, A&A, 542, L18
Lagrange, A.-M., et al., 2012b, A&A, 542, L40
Leggett, S., et al., 2010, ApJ, 710, 1627
Lowrance, P., et al., 1999, ApJ, 512, L69
Lowrance, P., et al., 2000, ApJ, 541, 390
Luhman, K. L., et al., 2007, ApJ, 654, 570
13
TABLE 1
Observing Log
UT Date
Telescope/Instrument Mode
Pixel Scale
(mas pixel−1)
Filter
tint Nimages
(s)
∆PA
(degrees)
2011-12-16
2012-01-11
2012-12-15
2012-12-16
2012-12-16
2012-12-23
2012-12-26
2013-01-09
VLT/NaCo
VLT/NaCo
Direct
Direct
13.22
13.22
VLT/NaCo
VLT/NaCo
VLT/NaCo
Gemini/NICI
Gemini/NICI
Gemini/NICI
Direct
Direct
Direct
Direct
Direct
0.′′22 mask
27.1
27.1
27.1
17.97
17.97
17.97/17.94
J
H
M ′
L′
[4.05]
[3.09]
[3.09]
H/Ks
50
40
20
30
30
38
38
11.4
48
92
176
112
64
60
60
117
26
34.6
70.2
67.7
57.4
30.2
31.7
41.1
β Pictoris b Detections and Photometry
TABLE 2
UT Date
Telescope/Instrument
Filter Wavelength (µm)
SNR
Apparent Magnitude
Absolute Magnitude
2011-12-16 VLT/NaCo
2012-01-01 VLT/NaCo
2013-01-09 Gemini/NICI
2013-01-09 Gemini/NICI
2012-12-23 Gemini/NICI
2012-12-26 Gemini/NICI
2012-12-16 VLT/NaCo
2012-12-16 VLT/NaCo
2012-12-15 VLT/NaCo
J
H
H
Ks
[3.09]
[3.09]
L′
[4.05]
M′
1.25
1.65
1.65
2.16
3.09
3.09
3.8
4.05
4.78
9.2
30
6.4
10
4.6
11
40
20
22
14.11 ± 0.21
13.32 ± 0.14
13.25 ± 0.18
12.47 ± 0.13
11.71 ± 0.27
--
11.24 ± 0.08
11.04 ± 0.08
10.96 ± 0.13
12.68 ± 0.21
11.89 ± 0.14
11.82 ± 0.18
11.04 ± 0.13
10.28 ± 0.27
--
9.81 ± 0.08
9.61 ± 0.08
9.54 ± 0.13
Sample Photometric NaCo and NICI Error Budgets
TABLE 3
Telescope/Instrument Filter Apparent Magnitude
Gemini/NICI
Gemini/NICI
VLT/NaCo
VLT/NaCo
Ks
[3.09]
L′
M ′
12.47 ± 0.13
11.71 ± 0.27
11.24 ± 0.08
10.96 ± 0.13
σdet
0.11
0.24
0.03
0.05
σatten
σf luxcal
0.06
0.11
0.07
0.11
0.03
0.08
0.03
0.06
Note. -- The photometric uncertainty from the intrinsic SNR (σdet) scales as
1.0857/SNR.
Luhman, K. L., 2013, ApJ, 767, L1
Macintosh, B., et al., 2008, SPIE, 7015, 31
Madhusudhan, N., Burrows, A., Currie, T., 2011, ApJ, 737, 34
Marois, C., et al., 2006, ApJ, 641, 556
Marois, C., et al., 2008, Science, 322, 1348
Marois, C., et al., 2010a, Nature, 468, 1080
Marois, C., et al., 2010b, SPIE, 7736, 52
Martinache, F., Guyon, O., 2009, SPIE, 7440, 20
Oberg, K., et al., 2011, ApJ, 743, L16
Pascucci, I., et al., 2006, ApJ, 651, 1177
Pecaut, M., et al., 2012, ApJ, 746, 154
Quanz, S., et al., 2010, ApJ, 722, L49
Rafikov, R., 2011, ApJ, 727, 86
Rameau, J., et al., 2013, A&A, 553, 60
Rayner, J., et al., 2009, ApJS, 185, 289
Rieke, G. H., et al., 2008, AJ, 135, 2245
Rodigas, T. J., et al., 2012, ApJ, 752, 57
Rousset, G., et al., 2003, SPIE 4839, 140
Skemer, A., et al., 2011, ApJ, 732, 107
Skemer, A., et al., 2012, ApJ, 753, 14
Soummer, R., Pueyo, L., Larkin, J., 2012, ApJ, 755, L28
Spiegel, D., Burrows, A., 2012, ApJ, 745, 174
Sudol, J., Haghighighpour, N., 2012, ApJ, 755, 38
Thalmann, C., et al., 2009, ApJ, 707, L123
Tokunaga, A., Vacca, W. D., 2005, PASP, 117, 421
Vigan, A., et al., 2012, A&A, 544, 9
Wahhaj, Z., et al., 2011, ApJ, 729, 139
Whitney, B., et al., 2003, ApJ, 598, 1079
Zuckerman, B., et al., 2001, ApJ, 562, L87
14
Young Directly Imaged Planets and Very Low-Mass Brown Dwarfs Used for Comparison
TABLE 4
Companion
D (pc)
Assoc.
Age
ST(Primary)
ST(Companion)
Sep.
Mass
References
Planets/Planet Candidates
HR 8799 b
HR 8799 c
HR 8799 d
HR 8799 e
κ And b
39.4 ± 1
39.4 ± 1
39.4 ± 1
39.4 ± 1
51.6 ± 0.5
Columba
Columba
Columba
Columba
Columba
30
30
30
30
30
A5
A5
A5
A5
B9IV
Low Mass
Brown Dwarfs
1RXJ 1609 B
GSC 06214 B
USco CTIO 108 B
HIP 78530 B
2M 1207 B
2M 1207 A
TWA 5B
HR 7329 B
PZ Tel B
2M0103AB B
AB Pic B
Luhman 16 B
Luhman 16 A
CD-35 2722 B
US
145 ± 14
US
145 ± 14
145 ± 14
US
156.7 ± 13.0 US
52.4 ± 1.1
52.5 ± 1.1
44.4 ± 4
47.7 ± 1.5
51.5 ± 2.6
47.2 ± 3.1
45.5 ± 1.8
2.02 ± 0.15
2.02 ± 0.15
21.3 ± 1.4
TWA
TWA
TWA
β Pic
β Pic
Tuc-Hor?
Carina
Argus?
Argus?
AB Dor
K7
5 -- 10
5 -- 10
K7
5 -- 10 M7
5 -- 10
∼8
∼8
∼ 8
12
12
30
30
40?
40?
∼ 100 M1Ve
B9V
M8
M8
M2Ve
A0
K0
M5/M6
K1Ve
L7.5
L7.5
??
L/T?
L/T?
L/T?
L2 -- L8?
L4 ± 2
L0 ± 1
M9.5
M8 ± 1
??
M8
M8 -- M8.5
M7.5
M7 ± 2
L?
L0±1
T0.5
L7.5
L4±1
67.5 -- 70.8
42.1 -- 44.4
26.4 -- 28.1
∼15
55± 2
4 -- 5
∼ 7
∼ 7
∼ 7
11.8-14.8
1,2,3,4,5
1,2,3,4,5
1,2,3,4,5
2,4,6
7
330
320 ± 30
670 ± 64
710 ± 60
40.8 ± 9
40.8 ± 9
∼ 98
200 ± 7
17.9 ± 0.9
84
248± 10
3.12 ± 0.25
3.12 ± 0.25
67.4±4
6 -- 12
14 ± 2
6-16
22 ± 4
8± 2
24± 6
∼ 20
26 ± 4
36 ± 6
12 -- 14
13 -- 14
40 -- 65
40 -- 65
31±8
8,9
10,11
12
11,13
14
14
15
16
17
18
19, 20
21,22
21,22
23
Note. -- References: 1) Marois et al. (2008), 2) Currie et al. (2011a), 3) Galicher et al. (2011), 4) Skemer et al. (2012), 5)
Currie et al. (2012b), 6) Marois et al. (2010a), 7) Carson et al. (2013), 8) Lafreni´ere et al. (2008), 9) Lafreni`ere et al. (2010),
10) Ireland et al. (2011), 11) Bailey et al. (2013), 12) Bejar et al. (2008), 13) Lafreni`ere et al. (2010), 14) Chauvin et al. (2004),
15) Lowrance et al. (1999), 16) Lowrance et al. (2000), 17) Biller et al. (2010), 18) Delorme et al. (2013), 19) Chauvin et al.
(2005), 20) Bonnefoy et al. (2010), 21) Luhman et al. (2013), 22) Burgasser et al. (2013), 23) Wahhaj et al. (2011)
Photometry for Young Directly Imaged Planets and Very Low-Mass Brown Dwarfs
TABLE 5
Companion
J
H
Ks
[3.09]
L′
M ′
Planets/Candidates
HR 8799 b
HR 8799 c
HR 8799 d
HR 8799 e
κ And b
Low-Mass
Brown Dwarfs
1RXJ 1609 B
GSC 06214 B
USco CTIO 108 B
HIP 78530 B
2M 1207 B
2M 1207 A
TWA 5B
HR 7329 B
PZ Tel B
2M0103AB B
AB Pic B
Luhman 16 B
Luhman 16 A
CD-35 2722 B
16.52±0.14
14.65±0.17
15.26±0.43
--
12.7 ± 0.30
15.08 ± 0.17
14.18 ± 0.17
14.23 ± 0.22
13.88 ± 0.20
11.7 ± 0.20
14.05 ± 0.08
13.13 ± 0.08
13.11 ± 0.08
12.89 ± 0.26
11.0 ± 0.4
--
--
--
--
--
12.68 ± 0.12
11.83 ± 0.07
11.50 ± 0.12
11.61 ± 0.12 > 10.09
9.54 ± 0.09
13.07 ± 0.30
12.05 ± 0.14
11.67 ± 0.35
--
12.09 ± 0.12
10.43 ± 0.04
10.72 ± 0.09
9.25 ± 0.05
16.40 ± 0.2
9.35 ± 0.03
9.1 ± 0.2
8.64 ± 0.19
8.70± 0.18
12.1 ± 0.3
12.80 ± 0.10
14.69 ± 0.04
15.00 ± 0.04
11.99 ± 0.18
11.06 ± 0.07
9.74 ± 0.04
9.94 ± 0.08
8.58 ± 0.04
14.49 ± 0.21
8.74 ± 0.03
8.65 ± 0.06
8.33 ± 0.1
8.31 ± 0.15
10.9 ± 0.2
11.31 ± 0.10
13.86 ± 0.04
13.84 ± 0.04
11.14 ± 0.19
10.38 ± 0.05
9.14 ± 0.05
9.30 ± 0.11
8.36 ± 0.04
13.31 ± 0.11
8.30 ± 0.03
7.91 ± 0.2
8.18 ± 0.1
7.86 ± 0.19
10.3 ± 0.2
10.76 ± 0.08
13.20 ± 0.09
12.91 ± 0.04
10.37 ± 0.16
9.84 ± 0.21
8.60 ± 0.08
--
--
--
--
--
--
--
--
--
--
--
--
8.99 ± 0.30
7.94 ± 0.07
--
7.99 ± 0.06
11.68 ± 0.14
7.73 ± 0.10
--
7.69 ± 0.1
--
9.3 ± 0.1
9.9 ± 0.1
--
--
--
--
7.94 ± 0.30
--
--
--
--
--
--
--
--
--
--
--
--
χ2
βP icb C.L.
52.8
6.098
8.351
--
0.946
∼ 1
0.893
0.961
--
0.186
0.287
0.864
--
1.369
7.001
--
88.086 ∼ 1
20.601 ∼ 1
48.044 ∼ 1
--
0.736
0.996
59.455 ∼ 1
--
2.666
11.231
--
--
--
Note. -- We only quantitatively compare the photometry between β Pic b and those objects with full JHKsL′
photometry.
15
TABLE 6
Adopted Flux Density
Zero-Points
Filter
λo (µm) Fν,o (Jy)
J
H
Ks
[3.09]
L′
[4.05]
M′
1.25
1.65
2.15
3.09
3.78
4.05
4.78
1594
1024
666.20
356
248
207
154
β Pictoris b Atmosphere Modeling Grid
TABLE 7
Model
Tef f
Limiting Cases
E60
1000-1800
AE60
1000-1700
A60
1000-1700
AMES-DUSTY 1000-2000
Range
log(g)
4 -- 4.5
4 -- 4.5
4 -- 4.5
3.5 -- 4.5
0.9 -- 2
0.9 -- 2
0.9 -- 2
0.9 -- 2
New Models
A4
1400-1900
3.6 -- 4.25
0.9 -- 2a
References
Rp (RJ )
1
2
1,3
4
5
Note. -- a) In our modeling, we perform two runs for the
A4 models: one where we fix the radius to values adopted
in the Burrows et al. (1997) hot-start models and one where
we allow the radius to freely vary between 0.9 RJ and 2 RJ .
References: 1) Burrows et al. (2006), 2) Madhusudhan et al.
(2011), 3) Currie et al. (2011a), 4) Allard et al. (2001), and 5)
this work.
TABLE 8
Model Fitting Results
Model
E60
AE60
A60
AMES-DUSTY
χ2
min
53.80
37.70
19.99
7.14
log(g), Tef f , R(RJ ) (for χ2
min)
log(g), Tef f , R(RJ ) (68%)
log(g), Tef f , R(RJ ) (95%)
4.0, 1400 K, 1.79
4.5, 1400 K, 1.96
4.5, 1400 K, 2.05
3.5, 1700 K, 1.35
--
--
--
--
--
--
--
3.5 -- 4, 1700 K, 1.32 -- 1.35
A4 (fixed radius)
5.85
3.8, 1600 K, 1.65
3.8, 1600 K, 1.65
3.6, 1500 -- 1550 K, 1.79 -- 1.80
3.8, 1550 -- 1625 K, 1.65
4.0, 1650 -- 1700 K, 1.54
A4 (scale)
A4 (1% CH4, scale)
.
5.82
5.13
4 -- 4.25, 1600 K, 1.65
3.6 -- 4.25, 1600 K, 1.64 -- 1.66
3.6 -- 4.25, 1500 -- 1750 K, 1.44 -- 1.82
4, 1575 K, 1.71
4, 1575 -- 1650 K, 1.59 -- 1.71
4, 1575 -- 1650 K, 1.59 -- 1.71
Note. -- The χ2 values quoted here do not refer to χ2 per degree of freedom. The columns for A4 (fixed radius) do not have a range in radius in some columns
because only one model (with a fixed radius) identifies the χ2 minimum
16
Fig. 1. -- Processed Gemini-NICI images obtained at H band (top-left), Ks band (top-right), and the H20 filter centered on ∼ 3.1 µm
(bottom panels: 23 Dec 2012 data on bottom-left, 26 Dec 2012 data on the bottom right). For clarity, we mask the region interior to ∼
0.′′4 and convolve the image with a gaussian equal to the image FWHM. The planet β Pic b is in the lower-right at a position angle of ∼
210o and a separation of ∼ 0.′′46. The color scale is set such that over the planet's FWHM the pixel color is white.
Fig. 2. -- Processed VLT/NaCo J (left) and H (right) band images presented in the same manner as Figure 1. Owing to very good
speckle suppression, the H-band data's effective inner working angle beyond which we are sensitive to β Pic b-brightness companions is
significantly smaller than for the J band data and the preceding NICI images (rIW A = 0.′′2).
17
Fig. 3. -- Processed VLT/NaCo L′ (left), [4.05] (middle), and M ′ (right) images presented as in Figures 1 and 2.
18
Fig. 4. -- Classical ADI reduction of our L′ data showing a clear detection of the β Pic debris disk. The green dot denotes the position
of β Pic b in the disk.
β Pic b
Planets/Low-Mass Brown Dwarfs
2.0
1.5
1.0
0.5
0.0
H
-
J
-0.5
-0.5
0.0
0.5
H-Ks
L
-
s
K
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
19
L
-
s
K
3.0
2.5
2.0
1.5
1.0
0.5
0.0
M Dwarfs
L0-L4 Dwarfs
L5-L9 Dwarfs
T Dwarfs
1.0
1.5
-1
0
1
J-Ks
2
3
K
-
J
2.5
2.0
1.5
1.0
0.5
0.0
-0.5
-1.0 -0.5
0.0
0.5
H-Ks
1.0
1.5
2.0
0.0
0.5
1.0
1.5
K-M
2.0
2.5
Fig. 5. -- Near-to-mid infrared color-color diagrams comparing β Pic b's colors (blue diamonds) to those of M dwarfs (small dark circles),
early L-type brown dwarfs (grey circles), late L dwarfs (asterisks), and T dwarfs (light grey dots) from Leggett et al. (2010). We also
overplot the positions of young, substellar objects/other directly-imaged planets (turquoise squares).
20
8
10
12
14
16
18
J
M
8
10
12
14
16
H
M
-1.0 -0.5
0.0
0.5
J-H
1.0
1.5
2.0
-0.5
0.0
0.5
H-Ks
1.0
1.5
8
10
12
14
16
H
M
9
10
11
12
13
14
15
K
M
0.5
1.0
1.5
2.0
2.5
3.0
0.0
0.5
1.0
H-L
1.5
K-M
2.0
2.5
Fig. 6. -- Color-magnitude diagrams comparing β Pic b to field M, L, and T dwarfs and young, substellar objects. Symbols are the same
as in Figure 5.
Fig. 7. -- Photometric data for β Pic b compared to L dwarf standard spectra between L2 and L8 as well as a low surface gravity L4
dwarf from Cushing et al. (2005, 2008). We scale β Pic b's flux density at Ks band to the band-integrated flux density of the standards at
2.15 µm. With the possible exception of the low surface gravity L4 dwarf, none of these standards provide a good match to measurments
for β Pic b. We identify the passbands along the bottom of the plot (left-hand panel).
21
Fig. 8. -- Comparisons between the β Pic b SED and the closest-matching substellar objects with J HKsL′ photometry. Quantitatively,
1RXJ 1609 B and κ And b provide the closest matches, although it is as-yet unclear whether any known substellar object fully reproduces
β Pic b's SED at all measured wavelengths.
22
Fig. 9. -- The χ2 distributions (top-left) and best-fitting models (top-right, middle panels) for our limiting cases: the E60, thin cloud/large
dust models the AE60, moderately-thick cloud/large dust models (second row), and the A60, thick cloud/large dust models. (Bottom
panels) The χ2 distributions and best-fit AMES-DUSTY models which assume ISM-sized dust grains. The horizontal lines in the lefthand
panels display the 68% and 95% confidence limits. The pink bars roughly denote the model-predicted flux densities at the filters' central
wavelength positions. We identify the passbands along the bottom of the plot (top-right panel).
23
Fig. 10. -- The effect of atmospheric dust particle sizes on the shape of a massive planet SED. Here, the < a > = 50, 30, and 5 µm models
depict log(g) = 4, Tef f = 1600 K models while the < a > = 3 µm model assumes log(g) = 3.8 and is scaled to match the luminosity of
the < a > = 5 µm model. For small particle sizes, surface gravity signatures are weak and this parameter's effect is primarily to change
the flux scaling.
Fig. 11. -- The χ2 distribution (left) and best-fitting model (right) for the thick cloud, small dust A4 models, showing that we can achieve
statistically significant fits to the data provided that the clouds are thick and the atmospheric dust particles are significantly smaller than
those we have previously assumed in matching L dwarf spectra (c.f. Burrows et al. 2006). This model fitting ties the planet radius to
predictions for hot-start models from Burrows et al. (1997). The horizontal dashed lines identify the 95% confidence limit (top) and 68%
confidence limit (bottom).
Fig. 12. -- Same as Figure 11 except treating the planet radius as a free parameter. Here we show the log(g) = 4, Tef f = 1600 K, though
the synthetic spectrum's appearance and its agreement with the data at neighboring gridpoints in surface gravity is nearly identical.
24
Fig. 13. -- SED fits adopting the nominally good-fitting atmosphere parameters depicted in Figure 12 but enhancing the atmosphere of
methane.
Fig. 14. -- (left) Luminosity evolution for hot-start models from Baraffe et al. (2003), comparing β Pic b's luminosity as derived in this
work to that for the directly-imaged planets HR 8799 bcde, the planet/low-mass brown dwarf κ And b, and other very low-mass brown
dwarf companions. The solid circle denotes β Pic b's nominal positions, whereas the open circle identifies its effective position on this plot
if it formed after 5 Myr. (Right) Evolution of the radius for planets with masses of 1 -- 10 MJ for the "hot-start" and "cold-start" planet
cooling models from Spiegel and Burrows (2012). The radius error bars define the range of planet radii from models consistent with the
data to within the 68% confidence limit.
|
1603.08930 | 1 | 1603 | 2016-03-29T20:00:05 | Probing Planet Forming Zones with Rare CO Isotopologues | [
"astro-ph.EP"
] | The gas near the midplanes of planet-forming protostellar disks remains largely unprobed by observations due to the high optical depth of commonly observed molecules such as CO and H$_2$O. However, rotational emission lines from rare molecules may have optical depths near unity in the vertical direction, so that the lines are strong enough to be detected, yet remain transparent enough to trace the disk midplane. Here we present a chemical model of an evolving T-Tauri disk and predict the optical depths of rotational transitions of $^{12}$C$^{16}$O, $^{13}$C$^{16}$O, $^{12}$C$^{17}$O and $^{12}$C$^{18}$O. The MRI-active disk is primarily heated by the central star due to the formation of the dead zone. CO does not freeze out in our modeled region within $70$AU around a sunlike star. However, the abundance of CO decreases because of the formation of complex organic molecules (COM), producing an effect that can be misinterpreted as the "snow line". These results are robust to variations in our assumptions about the evolution of the gas to dust ratio. The optical depths of low-order rotational lines of C$^{17}$O are around unity, making it possible to see into the disk midplane using C$^{17}$O. Combining observations with modeled C$^{17}$O$/$H$_2$ ratios, like those we provide, can yield estimates of protoplanetary disks' gas masses. | astro-ph.EP | astro-ph | Draft version June 11, 2021
Preprint typeset using LATEX style AASTeX6 v. 1.0
6
1
0
2
r
a
M
9
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
3
9
8
0
.
3
0
6
1
:
v
i
X
r
a
PROBING PLANET FORMING ZONES WITH RARE CO ISOTOPOLOGUES
Mo Yu 1, Karen Willacy 2, Sarah E. Dodson-Robinson 3, Neal J. Turner 2, and Neal J. Evans II 1
1Astronomy Department, University of Texas, 1 University Station C1400, Austin, TX 78712, USA
2Mail Stop 169-506, Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109
3University of Delaware, Department of Physics and Astronomy, 217 Sharp Lab, Newark, DE 19716
ABSTRACT
The gas near the midplanes of planet-forming protostellar disks remains largely unprobed by observa-
tions due to the high optical depth of commonly observed molecules such as CO and H2O. However,
rotational emission lines from rare molecules may have optical depths near unity in the vertical di-
rection, so that the lines are strong enough to be detected, yet remain transparent enough to trace
the disk midplane. Here we present a chemical model of an evolving T-Tauri disk and predict the
optical depths of rotational transitions of 12C16O, 13C16O, 12C17O and 12C18O. The MRI-active disk
is primarily heated by the central star due to the formation of the dead zone. CO does not freeze
out in our modeled region within 70 AU around a sunlike star. However, the abundance of CO de-
creases because of the formation of complex organic molecules (COM), producing an effect that can
be misinterpreted as the "snow line". These results are robust to variations in our assumptions about
the evolution of the gas to dust ratio. The optical depths of low-order rotational lines of C17O are
around unity, making it possible to see into the disk midplane using C17O. Combining observations
with modeled C17O/H2 ratios, like those we provide, can yield estimates of protoplanetary disks' gas
masses.
Keywords: astrochemistry -- planets and satellites:
formation -- protoplanetary disks -- ISM:
molecules
1. INTRODUCTION
The mass and surface density of the inner 30 AU of
protoplanetary disks are critical parameters that control
disk evolution and planet formation. Disk masses are
currently determined by (sub)millimeter observations of
dust, because the optical depth of continuum emission is
low at radii much larger than 10 AU (Williams & Cieza
2011). However, observations begin to lose sensitivity to
grains as they grow beyond the observing wavelength;
millimeter wave observations may thus underestimate
solid masses in the inner disk, where dust settling and
higher densities lead to more rapid growth (P´erez et al.
2012).
On the other hand, commonly observed molecular
lines often suffer from high optical depth, and are there-
fore unable to probe gas near the midplane. Molecules
that are important mass contributors to outer Solar Sys-
tem objects, such as H2O and NH3, freeze out very close
to the central star from the perspective of observers >
100 pc away. Thus, gas phase lines are hard to ob-
serve and most do not probe the giant planet-forming
region. Here we seek a molecule that can be used to
probe the conditions in planet-forming midplanes with
rotational transitions observable from the ground.
It
must be present in the gas phase, even in a cold disk,
and be optically thin enough that the disk is transpar-
ent in its rotational line emission. CO is volatile enough
not to freeze out within a few tens of AU of the central
star, so is able to probe a large disk area. Ali-Dib et al.
(2014b) even find that, like H2O, CO may be an im-
portant planet-building molecule for ice giants. While
rotational emission from 12C16O is optically thick due to
the molecule's high abundance, rare isotopologues of CO
may provide optically thin lines due to their low abun-
dances (Miotello et al. 2014).
If so, CO isotopologue
observations could lay the foundation for empirically de-
termining the mass available for gas giant formation in
nearby disks.
A clue that overall disk masses may be much larger
than the MMSN (minimum-mass solar nebula) came
from Bergin et al. (2013), who measured the mass of TW
Hya using the optically thin J = 1 → 0 transition of HD
with the Herschel Space Observatory. The 10 Myr-old
transitional disk is estimated to have a mass of 0.05 M(cid:12),
much heavier than the 0.01M(cid:12) of the MMSN, and larger
than many measurements estimated from dust emission.
This study has demonstrated the importance of measur-
ing gas mass directly. However, the HD J = 1 state has
a high excitation energy of E/kb = 128.5 K (kb is the
Boltzmann constant), so the line intensity depends on
not only the mass available but also the temperature of
the gas. Moreover, this transition is not accessible from
the ground. Since there is no far-infrared space mission
available, this method cannot be generally applied in the
near future.
Extrapolating from line intensities to total disk mass
requires knowledge of both the abundances and the iso-
topic ratios in the molecules being observed. Abun-
dances are not constant throughout the disk, as freeze-
out rate, UV intensity, and X-ray intensity all change
with distance from the star. Some chemical reactions
are isotope-selective, so the isotope ratio in molecules
is also a function of distance from the star. Miotello
et al. (2014) have shown that the abundance ratio of
rare isotopologues of CO to 12C16O (If not specified, CO
means the most common isotopologue of CO - 12C16O -
hereafter) can deviate from the atomic ratio significantly
due to isotope-selective photodissociation; the extent of
the effect depends on the strength of UV radiation and
dust properties. Finally, disk temperature and density
change over time, bringing changes in the molecular in-
ventory as the disk evolves. Here we present chemical
evolution models of protostellar disks to build a frame-
work for translating observed fluxes into mass densities
available for planet formation.
The disk model consists of three main components --
the thermodynamic model, the chemical model, and the
optical depth estimation. The workflow of the disk
model is sketched in Fig. 1, and the structure of this
paper follows the model workflow. The thermodynamic
model (Section 2) calculates the evolution of the disk
thermal structure for 3 Myr by adding the passive heat-
ing from the stellar irradiation to the viscous heating
from the magnetorotational
instability (MRI) turbu-
lence (Balbus & Hawley 1991). We construct 2D dust
radiative transfer models with RADMC (Dullemond &
Dominik 2004) to calculate the passive heating, and
adopt the viscous heating profile from an MRI-active
disk model (Landry et al. 2013), which includes a vis-
cosity prescription for accretion driven by MRI turbu-
lence. The chemical model (Section 3) then uses the den-
sity and temperature profiles from the thermodynamic
model as input to calculate the disk's chemical evolu-
tion for 3 Myr. After analyzing chemical model results
and discussing the model dependencies on initial molec-
ular cloud abundances, uncertainties in reaction rates,
and different grain evolution scenarios in section 4, we
estimate optical depths of various rotational emission
lines from CO isotopologues (Section 5), and calculate
the optical depths contributed by dust at correspond-
ing locations. Finally, we summarize the main results in
section 6. A list of symbols and definitions used can be
found in Table 1.
2. THE THERMODYNAMIC MODEL
The thermodynamic model provides density and tem-
perature evolution profiles for both the chemical model
and the optical depth calculation. Our thermodynamic
model is built upon the 1 + 1D disk model by Landry
et al. (2013), who calculated the structure and evolu-
tion of a disk with 0.015 M(cid:12) within 70 AU of the star.
The central star follows the evolutionary track of a
0.95 M(cid:12) star (D'Antona & Mazzitelli 1994) from 0.1 Myr
(T(cid:63) = 4600K, and R(cid:63) = 5.5 R(cid:12)), roughly the begin-
ning of the T-Tauri phase (Dunham & Vorobyov 2012),
to 3 Myr (T(cid:63) = 4500K, and R(cid:63) = 1.5 R(cid:12)). Landry
et al. (2013) followed the viscous evolution of the inner
70 AU of the disk. Because our goal is to probe the giant
planet-forming regions, we focus our study on the inner
70 AU of the disk.
The disk is heated passively by stellar irradiation, and
actively by viscous heating due to accretion. In order to
determine the degree of viscous heating, Landry et al.
(2013) evaluated whether MRI is present or not in each
individual grid cell by considering both Ohmic resistivi-
ties and ambipolar diffusion. The MRI is shut down by
Ohmic resistivity on the midplane, forming a deadzone
extending to about 16 AU. The disk has an MRI ac-
tive layer near the surface, which provides an accretion
rate of 10−9 M(cid:12)yr−1 even when a dead zone is present.
However, unlike models with assumed uniform turbulent
efficiency [with a constant α parameter, where viscosity
ν = αcsH, cs is the sound speed, and H is the scale
height; Shakura & Sunyaev (1973)], heating contributed
by MRI (extracted from the energy of shear flow) only
contributes slightly to the total heat budget of the disk
because the power is deposited in the disk atmosphere
at low optical depths.
Landry et al. (2013) calculated the heating from stel-
lar irradiation in a 1 + 1D geometry using Rosseland
mean opacities. Following Chiang & Goldreich (1997),
Landry et al. (2013) considered stellar radiation enter-
ing the disk at a grazing angle, heating up the disk sur-
face. Reprocessed stellar flux then traveled vertically
from the disk surface to the midplane. However, long-
wavelength radiation emitted by dust grains at the disk
surface should travel unimpeded to the midplane, caus-
ing heating that is not captured by using Rosseland
mean opacities. By assuming that heating propagates
only vertically from the heated surface, one also neglects
the heat transfer in the radial direction within the disk,
which can be important when the wavelength of the re-
emitted light is long enough to allow radial propagation.
One therefore underestimates the disk temperature by
using a 1 + 1D geometry and Rosseland mean opacities.
The stellar irradiation cannot heat the disk enough
2
to ionize the disk interior, so the 1 + 1D approximation
with Rosseland mean opacity in Landry et al. (2013)
was useful to calculate ionization, making the disk evo-
lution model tractable. However, the missing heating
can be crucial for the disk chemistry. We improved the
treatment of the stellar irradiation by including 2D dust
radiative transfer models built with the publicly avail-
able code RADMC (Dullemond & Dominik 2004). We
then calculated the total temperature by taking a flux-
weighted sum of the accretion temperature (Tacc) cal-
culated from the MRI model and the equilibrium tem-
perature (Teq) from stellar irradiation calculated from
the 2D dust radiative transfer model at every point of
the disk (as demonstrated in the upper half of Fig. 1.):
T 4 = T 4
eq + T 4
acc,
(1)
The 2D dust radiative transfer models are set up with
gas density profiles calculated by Landry et al. (2013)
and a constant gas/solid ratio throughout the disk. The
main focus of Landry et al. (2013) was to study the mass
transport. The disk is truncated at about two scale
heights above the midplane. The model thus includes
the majority of the disk mass, but the dust above the
original disk surface in Landry et al. (2013) could affect
disk thermal properties by absorbing the stellar radia-
tion. We therefore extrapolate the density profile verti-
cally with Gaussian distributions down to a background
density of 10−26 g cm−3 as input for the radiative trans-
fer calculation.
Landry et al. (2013) found that if the mean grain size
is as small as 0.1µm, the MRI shuts down completely.
If the MRI drives accretion in the T-Tauri phase, some
grain growth must have occurred, consistent with the
findings of Oliveira et al. (2010).
(See Gressel et al.
(2015) for updated models suggesting that accretion is
driven by winds, not MRI.) Evidence for rapid growth
of solids in disks has been accumulating in the follow-
ing ways: direct evidence from long-wavelength observa-
tions Isella et al. (2010), Guilloteau et al. (2011), Ban-
zatti et al. (2011), P´erez et al. (2012), P´erez et al. (2015),
and Tazzari et al. (2015); observations of forming plan-
ets in the LkHa 15 disk by Kraus & Ireland (2012); and
the HL Tau images from the ALMA Partnership et al.
(2015), which show gaps possibly sculpted by forming
planets or resonances with planets.
The wavelength-dependent dust opacities that we use
are taken from the website1, which provides models
whose Rosseland and Planck mean opacities are de-
scribed by Semenov et al. (2003), who updated earlier
1 http://www.mpia.de/homes/henning/Dust opacities/Opaciti-
es/opacities.html
3
models by Henning & Stognienko (1996) for dust in pro-
toplanetary disks. The available dust models include
those with different assumptions about the iron mix-
ture in the silicates, different models of grain growth,
and different grain topologies. The dust grains are ag-
gregated from "sub-grains", which themselves follow an
MRN distribution extended up to 5 µm (Mathis et al.
1977; Pollack et al. 1985). The dust model that we use
assumes a "normal" mix of iron [Fe/(Fe+Mg) = 0.3] and
that grain growth has occurred through particle clus-
ter aggregation (PCA in the nomenclature of Semenov),
which leads to roughly spherical grains. We use the mul-
tishell spheres topology and take the models for temper-
atures up to 155 K, in which ices and volatile organics
are retained. Figure 2 shows the opacity per gram of
gas versus wavelength for the adopted dust properties.
There are a number of resonances and a relatively slowly
declining opacity out to a wavelength of about 1.5 mm,
beyond which the opacity declines rapidly. The opacities
are given per gram of gas.
In our standard model, the dust has already grown
and aggregated at the start of disk evolution - 0.1 Myr
after the formation of the central star. We further as-
sume that 90% of the dust has grown to still larger sizes
(pebbles, rocks, etc.), which we no longer describe as
dust. This further growth decreases the grain surface
area for chemical reactions and makes these solids es-
sentially invisible even to millimeter wavelength obser-
vations (Birnstiel et al. 2011; Garaud et al. 2013). Con-
sequently, we take a gas to dust ratio of 1000, so we
divide the opacities from Semenov et al. (2003) by 10.
In this model, there is a constant replenishment of dust
by collisions between larger objects and the size distri-
bution of dust does not evolve. We consider a different
model for grain evolution as a variation of our standard
assumptions in §4.5.3.
The temperature color coded plots (at the beginning
and end of the evolution) with Teq computed from the
2D dust radiative transfer model are shown in the up-
per panels of Fig. 3. In both cases, the temperature is
the highest on the disk surface due to the heating from
the central star, and gradually decreases at larger radius
and into the disk interior. The temperature almost ev-
erywhere in the disk decreases due to the star evolving
down the Hayashi track (L(cid:63) decreases from 12.1 L(cid:12) to
0.8 L(cid:12)) and the flattening of the disk itself (cf. the top
two panels in Fig. 3). The accretion heating (Tacc) con-
tributed by the MRI turbulence is shown in the lower
left panel of Fig. 3. Accretion heating contributes less
then 3 K in the majority of the disk, rising to about 4 K
within 2 AU from the central star.
We compare the midplane temperature with Teq com-
puted from the 1 + 1D model after 3Myr years of evolu-
tion in Landry et al. (2013) and the one computed from
the 2D dust radiative transfer model in the lower right
panel of Fig. 3. The temperature is much higher in the
2D model in the inner 20 AU. The 1 + 1D model misses
heating in the inner part of the disk where the optical
depth is larger, and therefore produces artificially lower
temperature on the midplane closer to the central star.
For example, the Rosseland mean optical depth is ∼ 12
at 5 AU, but ∼ 4 at 15 AU, so 5 AU is more severely
affected -- even though its surface temperature is higher.
Moreover, once the temperature is down to ∼ 50 K, the
black-body radiation re-emitted from dust peaks at far-
infrared wavelengths. The mean free path is a significant
fraction of the scale height and the assumptions used to
justify using Rosseland mean opacity breaks down.
Our treatment of temperature is not ideal in the sense
that the vertical density structure and the temperature
structure are not calculated consistently. However, as
an experiment we artificially increased the scale height
everywhere in the disk by 20 percent and did not find
significant changes to the midplane temperature. We
also assume the gas and dust are well mixed and have
the same temperature in this paper, which is a valid as-
sumption except for the very surface of the disk. Since
the purpose of this project is to find an optical depth in
the vertical direction and we mainly focus on the disk
midplane where most of the disk mass resides, the com-
bined accretion and passive heating models, assuming
the same temperature for the gas and dust, meet our
needs.
In the next sections, we describe our standard model
and results. Then we describe two variations in the
chemistry and one variation in the model of dust evolu-
tion.
3. CHEMICAL MODEL
The disk temperature and density evolution described
in Section 2 allow us to compute the chemical evolution
of the disk.
We also need initial chemical abundances, which we
obtain from a simplified model of the molecular cloud.
We construct chemical evolution models including C,
H, O, N, and different C and O isotopes based on the
UMIST database RATE06 (Woodall et al. 2007), and
follow the chemistry of 588 species, 414 gas-phase and
174 ices. The reaction network contains 13116 reactions,
including gas-phase reactions, grain-surface reactions,
freezeout, thermal desorption, and reactions triggered
by UV, X-rays and cosmic rays, such as isotope-selective
photodissociation. The carbon isotopic chemistry net-
work was developed by Woods & Willacy (2009) and
was extended to include oxygen isotopes for this work.
We calculate the chemical evolution at each grid point
independently. By doing this, we assume that the radial
and vertical motions of gas and dust are slow compared
4
to chemical reaction timescales, and that mixing is not
important in determining the chemistry. This is a valid
assumption for gas-grain reactions or grain surface re-
actions due to their short reaction timescale, but might
not work as well for gas-phase reactions. However, given
that turbulence in the inner 15 AU is restricted to the
disk surface, we do not expect vertical mixing to con-
tribute much to disk chemistry.
In order to calculate the chemical evolution over a
3 Myr lifetime, we use an extension of the computa-
tionally efficient rate-equation method to compute re-
action rates for grain-surface reactions under the "mean
field" approximation (as used by Dodson-Robinson et al.
2009). Rate equations neglect the stochastic variation
of abundances on different grains and are not appropri-
ate when the number of reacting particles per grain is
less than one. We modify our reaction rates following
the method of Garrod & Pauly (2011), which consid-
ers the competition between thermal hopping of mobile
species and reactions on the grain surfaces. We assume
that only atoms and simple hydrides are mobile on the
grains. Our treatment of gas phase reactions, freezeout,
and thermal desorption is similar to that of Dodson-
Robinson et al. (2009). After describing the molecular
cloud preprocessing model in section 3.1, we focus the
rest of this section on the treatment of photochemical
and cosmic-ray reactions, which are new to this work.
3.1. Preprocessing in the molecular cloud
We model the chemical evolution in the molecular
cloud stage to derive input abundances for the proto-
planetary disk models.
The input abundances are listed in Table 2 (Graedel
et al. 1982). Throughout this paper, abundances are
presented as the number density normalized to the num-
ber density of hydrogen nuclei, nx/(nH + 2nH2 ).
The molecular cloud model is run for 1 Myr with a
density of 2 × 104 cm−3, at a temperature of 10K and
with a visual extinction of 10 magnitudes.
The abundances at the end of the molecular cloud
phase are given in Table 4. The 12C/13C is ∼ 46 in
CO gas, ∼ 63 in CO ice and ∼ 58 in CO2 ice at the
end of the molecular cloud model -- all lower than the
initial elemental 12C/13C ratio. The reduction is caused
by the preferential formation of heavier isotopomers of
CO arising from the small energy different in the ion
exchange reaction.
13C+ + CO ↔13 CO + C+ + 35K.
(2)
13CO has a lower ground-state vibrational energy due
to its slightly larger mass compared to CO, and is there-
fore more energetically favorable. At 10 K Equation
2 dominates the carbon fractionation, leading to lower
12CO/13CO than the elemental 12C/13C ratio (see also
Visser et al. 2009, Langer & Penzias 1993).
In other
molecules, such as CH4 ice, the opposite effect is seen,
with enhanced abundances of condensed 12C leading to
higher 12CH4/13CH4 ratios in the ice.
3.2. UV ionization, photodissociation and
photodesorption
In addition to ionizing and dissociating gas phase
species, UV photons also desorb molecules from grain
mantles. The UV flux generated by a young star has
been observed by France et al. (2014) who found a me-
dian value of 1000 ISRF (1 ISRF = 1.6 × 10−3 ergs−1
cm−2; Habing (1968)) at 100 AU. Theoretical studies of
Alexander et al. (2006) suggest a UV flux of ∼ 50 ISRF
at 100 AU is required to drive photoevaporation of the
disk. We choose a UV flux between these two values of
500 ISRF at 100 AU.
The UV field is attenuated by the disk and the re-
sulting visual extinction can be related to the column
density by
AV (r, z) = 5.2 × 10−22 × fH Σr(r, z)
2mH
,
(3)
where fH = 0.735 is the mass fraction of hydrogen in
the Sun (Grevesse & Sauval 1998), and Σr(r, z) is the
horizontal column density integrated from the inner edge
of the disk.
For photodesorption we adopt the reaction rate of Hol-
lenbach et al. (2009)
k(r, z) = F (r, z) × 108 × (πa2)×Y,
(4)
where F(r,z) is the UV flux in ISRF, πa2 is the grain
cross-section and Y is the photodesorption yield. The
average grain radius a is taken to be 1 µm, and Y is as-
sumed to be 10−3 for all molecules (based on laboratory
measurements of H2O photodesorption by Lyα photons
(Westley et al. 1995a, b).
For the self-shielding of CO and H2 we use the method
of van Dishoeck & Black (1988) and Lee et al. (1996)
respectively.
3.3. X-ray photoionization
We follow the X-ray ionization prescription of Bai &
Goodman (2009), using a value of the ionization rate
for direct absorption of X-rays, ζabs = 6 x 10−12 s−1,
and for scattered photons, ζsca = 10−15 s−1 (Igea &
Glassgold 1999). We assume a stellar X-ray luminos-
ity LX = 20 × 1029 erg s−1 based on the median ob-
served value from a Chandra survey of solar mass young
stars in the Orion Nebula (Garmire et al. 2000). Simi-
lar values have also been observed in the Taurus-Auriga
complex by Telleschi et al. (2007) and Robrade et al.
(2014). The X-ray ionization rate is assumed to be the
same for all reactions because of the lack of laboratory
measurements.
3.4. Cosmic Rays
Cosmic rays can ionize and dissociate molecules in the
gas directly, or by producing secondary photons (Gredel
et al. 1989). The intensity of cosmic rays decreases ex-
ponentially with a characteristic attenuation depth of 96
g cm−2 (Umebayashi & Nakano 1981, 2009). The large
penetration depth means that our model disk is trans-
parent to cosmic rays outside of 4 AU at the beginning
of the T Tauri phase and outside of 3 AU at the end
of 3 Myr evolution. Cosmic ray reaction rates are taken
from UMIST06 with an assumed cosmic ray ionization
rate of 1.3 × 10−17 s−1.
3.5. A summary of ionization processes
Ionization rates contributed by the above three ioniz-
ing sources are shown in Fig. 4. The upper left panel
shows an order-of-magnitude estimate of the UV ion-
ization rate based on the H2O molecule, which is the
molecule most commonly ionized by UV. The ionization
rate decreases very rapidly due to the small penetra-
tion depth of UV photons, and it is negligible except for
the very surface layer of the disk. X-rays are able to
reach most of the disk except for the inner 10 AU near
the disk midplane (shown in the upper right panel in
Fig. 4). They provide ionization rates of 10−14 s−1 to
10−13 s−1 near the disk surface, and a modest ioniza-
tion rate around 10−17 s−1 in the disk interior due to
scattering. In the lower left panel, we show the cosmic-
ray ionization rate of H2, the most common cosmic-
ray reaction partner. Cosmic rays provide a steady
ionization rate around 10−17 s−1 throughout the disk.
The fractional contribution of X-ray ionization is plot-
ted in the lower right panel of Fig. 4. Roughly speak-
ing, X-rays dominate the ionization rate above one scale
height of the disk where the vertical column density is
less than a few g cm−2, and the cosmic ray ionization
dominates the disk interior within one scale height of
the midplane. Cosmic ray-induced photons are abun-
dant enough to cause ionization and photodissociation
throughout the disk due to the efficient penetration of
cosmic rays. Ionization rates due to cosmic ray-induced
photons are different for each reactant and are not shown
in Fig. 4. However, cosmic ray-induced photons are im-
portant contributors to the chemical evolution.
Cleeves et al. (2013a) showed that stellar winds can
power a "T-Tauriosphere" that shields the disk from ex-
ternal cosmic rays, leading to cosmic ray ionization rates
of 10−18 s−1 or lower. A decreased cosmic-ray flux would
reduce the rates of reactions with both cosmic rays and
cosmic ray-induced photons. On the other hand, the
decay of shortlived radionuclides (SLRs) such as 26Al
5
can provide an ionization rate on the order of 10−19 s−1
to 10−18 s−1 (Cleeves et al. 2013b), which may help to
drive the chemistry in the disk interior. Since our chem-
ical evolution models are computationally expensive, we
do not explore different values of ionization rates.
4. CHEMICAL MODEL RESULTS
The results of our chemical model give the abundances
of CO isotopologues that we need for the optical depth
calculation. First, we discuss the ice line locations for
different volatiles (Section 4.1). Knowing the ice line
locations is important for interpreting observed radial
abundance gradients. Second, we discuss the active
reaction network involving carbon-bearing molecules,
which causes the CO abundance to change with both
radius and time (Section 4.2). In Section 4.3, we demon-
strate that dissociation of CO and subsequent forma-
tion of complex organic molecules (COM) can produce
CO depletion that mimics an ice line.
In Section 4.4,
we assess how computational limits on the number of
species in our reaction network impact our COM abun-
dances. We discuss our model's dependence on initial
cloud abundances, reaction rates, and grain evolution
in Section 4.5. We show that despite the uncertainty in
the exact end product of COM formation, the formation
of ices as carbon sinks on grain surfaces is robust, and
leads to the depletion of CO in the gas phase. In order to
connect the abundances of rare CO isotopologues with
the overall disk mass, we examine the carbon fraction-
ation in CO and other major carbon-bearing molecules
in section 4.6.
4.1. Locations of ice lines
Due to efficient heating from the central star, CO and
CH4 do not freeze out in our modeled region -- the inner
70 AU of the disk -- at any time in our 3 Myr of evolution
(though disks with different grain properties or less lumi-
nous host stars may be colder than our model disk). Us-
ing CO isotopologues to estimate planet-forming mass
therefore does not require correcting for CO freezeout, at
least in disks surrounding proto-Sunlike stars. However,
as we will see in Section 4.3, freezeout of other organic
molecules affects the gaseous CO abundance. H2O and
CO2 freeze out at 1.5 AU and 18 AU at the beginning
of disk evolution, and their condensation fronts move
inward to 1 AU and 2 AU as the disk becomes cooler.
We see ices of hydrocarbons and other carbon-bearing
molecules formed from CO after a few hundred thou-
sand years of disk evolution. The abundance of an ice
depends on both condensation temperature and forma-
tion pathway: a molecule may have been chemically de-
stroyed and simply not be present to freeze out. Our
discussion of ice lines focuses only on locations where
the relevant molecule exists. We defer the discussion
of molecule formation and destruction to the next two
sections.
The binding energies and the locations where different
ice species exist after 3 Myr of disk evolution are shown
in Table 3. Species with larger binding energy can form
stronger bonds with the grain surface, and therefore
freeze out at higher temperature. In our model, species
with binding energy larger than EB/k = 2.5 × 103 K
can freeze out well within the giant planet-forming re-
gion at (cid:46) 15 AU (Tsiganis et al. 2005; Thommes et al.
2002); species with binding energy of 2.1 × 103 K can
only freeze out in the outer part of the disk or at later
stage of evolution when the temperature is lower, and
species with binding energy lower than 103 K -- including
CO -- do not freeze out.
The condensation of volatiles is important for the
growth of planets due to the increase of available solid
surface density (Dodson-Robinson & Bodenheimer 2010;
Ali-Dib et al. 2014b), and is crucial for determining
the chemical composition of giant planets ( Oberg et al.
2011b). Moreover, the effect of radial drift and gas ac-
cretion may cause further movement of ice lines and
changes in chemical composition (Ali-Dib et al. 2014a;
Piso et al. 2015). However, the above studies assumed
CO to be a major carrier of carbon and oxygen, which
may not be true throughout the disk evolution. Our re-
sults indicate that it is important to take into account
the possibility that CO ice is not a mass source for giant
planets in some systems. Instead, CO2, hydrocarbons,
and methanol may be the major carbon ice reservoirs
that contribute to planet growth.
Furthermore, because of the unprecedented sensitiv-
ity and resolving power of (sub)mm interferometers such
as the Atacama Large Millimeter/submillimeter Array
(ALMA), the location of the CO ice line has been con-
sidered an important temperature tracer of giant planet-
forming regions (Qi et al. 2011, 2013). Qi et al. (2013)
use observations of N2H+, which CO destroys, to infer a
CO snow line radius of ∼ 30 AU in the disk surrounding
TW Hya. If some combinations of grain size, star lumi-
nosity and UV/X-ray/cosmic-ray flux push CO ice lines
outside of 70 AU, CO freezeout may not trace the region
of giant planet formation. Even TW Hya, at 0.8 M(cid:12) and
10 Myr, is luminous enough to push its disk's CO snow
line beyond the likely formation locations of Uranus and
Neptune in the solar nebula.
4.2. Time evolution and spatial distribution of CO gas
Inferring disk structure based on CO line intensity
is complicated by the fact that the abundance of CO
evolves significantly as a function of time.
In the in-
ner 20 AU from the central star when the temperature
is high enough for CO2 to be in the gas phase, CO2 is
dissociated into CO + O by cosmic ray-induced photons
6
at a rate around 10−18 per second, which leads to an in-
crease of CO abundance on a 1 Myr time scale. On the
other hand, although CO does not freeze out in our mod-
eled region within 70 AU, the abundance of CO drops
beyond 15 AU because carbon is tied up in hydrocar-
bons, methanol, and ketene (complex organic molecules
or COMs), mimicking the effect of CO freezeout. Abun-
dant carbon-bearing species include C2H2 (acetylene),
C2H5, CH3CHO (acetaldehyde), CH3OH (methanol),
and H2CCO (ketene). The CO depletion rate is driven
by the ionization rate of He+ from X-rays and cosmic
rays, as He+ drives the breakup of CO (section 4.3). The
depletion of CO beyond 15 AU happens on a 1 Myr time
scale. As a result, the CO abundance depends strongly
both on location and time. As long as the disk is opaque
to the UV photons that dissociate CO, the column den-
sity of CO can increase in the inner part of the disk even
as the overall disk mass is decreasing.
Color-coded plots of abundances of major carbon-
bearing molecules are shown in Fig. 5 and Fig. 6. We
see abundant CO within 15 AU from the central star,
and abundant CO2 ice in parts of the disk where the
temperature is low enough for it to stay on the grain
surface. H2CCO (ketene) ice is found to be the major
carbon sink between 5 AU and 45AU from the central
star in our model. Other COMs such as C2Hx, CH3OH
(methanol) and CH3CHO (acetaldehyde) exist in a layer
between 2 − 30 AU, closer to the disk surface, or the
whole disk thickness beyond 40 AU from the central star
(lower right panel of Fig. 5). Again, the change of abun-
dance can be very gradual. Here we quote the boundary
location where 10% of carbon is stored in corresponding
species. At the end of our 3 Myr disk evolution, 13.6% of
available carbon is contained in CO gas, 36.2% in CO2
ice, and 44.8% in complex organic molecules. The above
values are integrated over the entire disk, weighted by
the disk mass in different locations. A detailed break-
down of abundances of various species can be found in
the first column of Table 5. We present abundances
in the form of percentage of carbon contained in each
species, integrated over the entire disk, and weighted by
disk mass in different locations.
The time evolution of the CO/H2 abundance ratio
(Fig. 7) over 3 Myr is substantial. The variation is not
well-represented by a step function as in simple freeze-
out models. Models of CO abundance versus radius ap-
propriate to the star's age must be used to compute
available planet-forming mass from CO isotopologue line
intensities.
While this paper does not contain a parameter study
of CO/H2 abundance ratio as a function of disk mass and
UV/X-ray/cosmic-ray flux, we suggest that flux in high-
velocity line wings (produced by CO gas near the star)
may increase as the star ages, even despite an overall
7
reduction in disk mass. Furthermore, depletion of CO
gas in the outer disk does not necessarily mean there is
CO frozen on grain surfaces, an idea we explore further
in the next section.
4.3. CO depletion due to the formation of complex
organic molecules
Here we investigate the causes of CO depletion in the
outer disk, beyond 15 AU. As we can see in Fig. 5 and
Fig. 6, H2CCO (ketene) ice is the dominant form of
carbon within 45 AU from the central star in the CO-
depleted region. Beyond 45 AU, CH3OH, C2Hx, and
CH3CHO ice are more abundant. To demonstrate the
chemical evolution in these two different locations, we
plot the abundances of major carbon-bearing molecules
as a function of time in Fig. 7 for two locations of the
disk: 38 AU on the midplane and 60 AU on the mid-
plane.
At 38 AU on the midplane, the depletion of CO hap-
pens in three stages. In stage 1, the first 0.6 Myr, CH4
and CO react to form C2H2 through two different paths
(Path 1 and Path 2, see the next paragraph for de-
tails), both of which depend on the existence of CH4
gas. This leads to a net destruction of CO and methane
and increase in C2H2 abundance. C2H2 freezes out on
grain surfaces, but because of the low binding energy
of C2H3, C2H2 can not efficiently hydrogenate until the
temperature decreases. The formation of more complex
organic molecules such as H2CCO and C2H5 happens
roughly between 0.6 to 1.5 Myr (Stage 2) and only after
the reactions in Stage 1 have already begun to deplete
CO. CH3OH ice forms in Stage 3 after the formation of
H2CCO and C2H5 ices. The CO gas abundance contin-
ues to decrease through all three stages. At 60 AU on the
midplane where temperature is lower, C2H3 can stay on
the grain surface and hydrogenate to C2H5 early in the
disk evolution. We see rapid C2H5 formation in the first
0.6 Myr, and the formation of hydrocarbons stops after
CH4 is depleted. Instead, the net transfer of carbon is
from CO to CH3OH.
The formation of COM primarily follows two paths:
reactions between CHx radicals (Path 1), and ion-
neutral reactions between C+ and CHx (Path 2). The
first pathway starts from CH4 dissociation by secondary
photons (hν) generated by cosmic ray ionization:
hν + CH4 → CH2 + H2.
(5)
CH2 then reacts to form CH and subsequently com-
bines with CH4 to form C2H4.
H + CH2 → CH + H2
CH + CH4 → C2H4 + H.
(6)
(7)
Similar reactions involving other CHx radicals include:
CH2 + CH2 → C2H3 + H
C + CH3 → C2H2 + H.
(8)
(9)
In a relatively high temperature and hydrogen rich
environment, C2H3 and C2H4 formed in above reactions
quickly react into a more stable form - C2H2.
C2H+
3 + C2H4 → C2H+
H + C2H3 → C2H2 + H2.
5 + C2H2
(10)
(11)
While the major carbon source for the first pathway is
CH4, this pathway also creates molecules that react with
byproducts of CO destruction. Since CO abundance is
our focus in this work, we direct our attention to the
second pathway. The reaction network that moves car-
bon from CO into COM is sketched in Fig. 8. The
formation of COM following CO destruction is primar-
ily driven by helium ionized by cosmic rays and X-rays.
Of the resulting He+ ions, more than half go on to dis-
sociate CO (creating C+ and O), while other He+ ions
end up ionizing molecules such as H2, C2H2, N2 and
CH4. C+ rapidly reacts with CH4 and CH3 to form
hydrocarbon ions. Hydrocarbon ions can go through
a series of charge-exchange reactions until they eventu-
ally recombine with an electron to form neutral hydro-
carbons. If the binding energy of the resulting neutral
molecule is large enough, the neutral will quickly freeze
onto the grain surface and remove carbon from the gas-
phase chemistry. Key reactions that initiate hydrocar-
bon formation are:
He+ + CO → O + C+ + He
C+ + CH4 → C2H+
3 + H
C+ + CH4 → C2H+
2 + H2
C+ + CH3 → C2H+ + H2.
reaction chain terminates
full
(12)
(13)
(14)
(15)
This
in frozen-out
methanol, acetaldehyde, and ketene sinks (see Fig. 8).
The electrons with which the C2-based hydrocarbon
ions eventually recombine come mostly from cosmic-ray
and X-ray ionization of H2. However, unlike He+, H+
2
does not contribute to hydrocarbon formation directly.
The majority of H+
2 initiates HCO+ formation through:
2 + H2 → H+
H+
3 + CO → HCO+ + H2.
H+
3 + H
(17)
(16)
Reactions involving HCO+ often change the charge
and/or saturation of a hydrocarbon (e.g.
C2H2+
HCO+ → CO + C2H+
3 ), but do not contribute to the
initial formation of the carbon-carbon bond.
After hydrocarbon ions recombine with electrons, the
8
resulting neutral molecules may freeze onto grain sur-
faces. The dissociative recombination reactions
C2H+
C2H+
C2H+
C2H+
C2H+
3 + e− → C2H + 2H
3 + e− → C2H2 + H
4 + e− → C2H2 + 2H
5 + e− → C2H2 + H2 + H
5 + e− → C2H3 + 2H
(18)
(19)
(20)
(21)
(22)
contribute the most to the total neutral C2Hx abun-
dance, and to the C2Hx ice budget (recombination path-
ways are marked with e− in figure 8). As a result, the
reaction of CO with He+ starts a chain that moves car-
bon atoms from CO to hydrocarbons on million-year
timescales.
4.4. Chemical pathways in the formation of complex
organic molecules
We have seen how the CO abundance beyond 20 AU
decreases over time due to complex organic molecule
(COM) formation. Here we trace the chemical path-
ways that transfer carbon from CO to COM and assess
whether the outer-disk CO depletion is robust given the
construction of our reaction network.
4.4.1. Ketene as a carbon sink
Our chemical model does not include any molecule
with more than two carbon atoms. Without more com-
plex species available for reaction outcomes, H2CCO
(ketene) is the most abundant two-carbon molecule
found in our model and serves as a "sink" for frozen-out
carbon on the grain surfaces (as seen in Fig. 5). The low
binding energy of C2H3 also contributes to the H2CCO
abundance by shutting off the hydrogenation pathway
from C2H3 to C2H6 on the grain surfaces and ensuring
that saturation must take place in the gas phase. (The
binding energies of abundant COMs are shown in Table
3.) Finally, the activation barrier of E/kB = 1210 K
for the reaction C2H2 + H → C2H3 (Hasegawa et al.,
1992), which has to break the strong C -- C triple bond,
means that at the cold temperatures required for acety-
lene to freeze onto grain surfaces, hydrogenation pro-
ceeds slowly.
The most common reaction for C2H3 in the gas phase
is
C2H3 + H → C2H2 + H2,
(23)
followed by C2H2 re-freezing onto grain surfaces. A
small fraction of C2H3 gas reacts with ions such as
C2H+
3 , HCO+, and forms more saturated hydrocarbons,
then refreezes onto dust grains:
3 → C2H+
C2H3 + C2H+
C2H3 + HCO+ → C2H+
5 + C2H,
4 + CO.
(24)
(25)
At the same time, a small fraction of C2H3 forms a
double bond with free oxygen released in Eq. 12:
C2H3 + O → H2CCO + H.
(26)
Since H2CCO has very high binding energy (only
slightly smaller than CO2), and we are not including de-
struction of H2CCO ice other than desorption, H2CCO
ice serves as a sink for COM in our model. Although
the rate for reaction 26 is low, ketene can neverthe-
less accumulate on million-year timescales. The reaction
pathway leading to ketene formation is sketched in the
lower left part of Fig. 8, with the sink molecule H2CCO
enclosed in a solid box. One should note that ketene
is a representative of the presence of complex organic
molecules in our model. In a real disk, other forms of
COM will likely exist.
Densities on grain surfaces are usually much higher
than in the gas phase, so reactions such as hydrogenation
on grain surfaces typically happen very quickly. How-
ever, due to the rapid thermal desorption of C2H3 from
the grain surface, more saturated hydrocarbons cannot
be formed efficiently. In most of the disk within 40 AU,
only a small amount of more saturated C2 hydrocarbons
can be formed in the gas phase through ion-neutral re-
actions, with a negligible amount formed though hydro-
genation on the grain surface.
4.4.2. Acetaldehyde as a carbon sink
Ketene is no longer the sink in regions where the tem-
perature is high enough for it to be in the gas phase,
or where the temperature is low enough for C2H3 to
stay on the grain surface and continue hydrogenating
to form more saturated hydrocarbons. In those regions,
ice molecules such as CH3CHO (acetaldehyde), CH3OH
(methanol), and C2H5 (ethyl) serve as the carbon sinks,
keeping carbon from re-entering the gas phase.
Green lines in Fig. 8 show the ketene recycling path-
way in a small region below the disk surface between
10 − 30 AU, where the temperature is high enough that
H2CCO stays in the gas phase and the ionization rate
is (cid:38) 10−15 s−1. H2CCO gas reacts with C+ to form
CH2CO+:
C+ + H2CCO → CH2CO+ + C.
(27)
CH2CO+ then recombines with an electron and under-
goes one of three dissociative recombination reactions
to form C2, C2H2, or CO, with roughly equal branching
ratios (see green lines leading from CH2CO molecule in
lower left corner of Fig. 8). Moreover, due to the high
ionization rate, the rate of reaction 12 can be as much
as twice the value on the disk midplane. The abun-
dance of free hydrogen atoms is also high due to pho-
todissociation of H2 and H2O. Efficient C+ production
and high atomic hydrogen abundance speed the produc-
tion of C2H2 and allow more saturated hydrocarbons to
form on grain surfaces, despite the volatility of C2H3.
The temperature in the region is high enough for C2H5
to evaporate into the gas phase once it is formed. C2H5
then reacts with atomic oxygen to from the more stable
molecule CH3CHO (acetaldehyde), which refreezes on
to the grain surface due to its high binding energy and
serves as another carbon sink:
O + C2H5 → CH3CHO + H.
(28)
To summarize, the acetaldehyde-forming reaction path-
way (Fig. 8) differs from the ketene-forming pathway
simply because of the warmer temperature that keeps
ketene in the gas phase, the high C+ abundance from
reaction 12, and the high atomic hydrogen abundance.
Note that our reaction network does not include the
grain-surface hydrogenation pathway C2H5 +H → C2H6
(ethane ice). However, Dodson-Robinson et al. (2009)
found low grain-surface hydrogenation efficiency, lead-
ing to significant amounts of ethane ice only near the
acetylene sublimation temperature of 55 K.
4.4.3. Ethyl and methanol as carbon sinks
In the outer disk where r > 50 AU, and after 1 Myr
in most of the disk, C2H3 is able to stay on the grain
surface and hydrogenate to C2H5, and C2H5 serves as a
carbon sink.
Due to the lack of C2H3 in the gas phase, and the
low temperature that allows H2CO to freeze out, the
reaction
O + CH3 → H2CO + H
(29)
becomes the dominant reaction with atomic oxygen,
rather than reaction 26. H2CO then freezes onto the
grain surface, reacts with H atoms on the grain surface,
and finally forms CH3OH (methanol), another carbon
sink in our model, as follows (letter G denotes species
that are frozen out on grain surfaces):
GH + GH2CO → GCH2OH
GH + GCH2OH → GCH3OH.
(30)
(31)
This reaction path explains the evolution of abun-
dances of carbon-bearing molecules demonstrated in the
lower panel of Figure 7. These reactions start to happen
at r > 60 AU where the temperature is about 35 K, and
at 20 AU on the midplane at 2.5 Myr when the temper-
ature drops below 44 K.
In summary, as Fig. 5 and Fig. 6 show, H2CCO
dominates in most of the COM-forming region where
the temperature is low enough for H2CCO to remain on
the grain surface, but high enough for C2H3 to evaporate
into the gas phase. In the small region above the disk
midplane (r = 10 − 35 AU, z = 1 − 4 AU), where the
temperature is high enough for H2CCO and C2H5 to
9
stay in the gas phase and atomic hydrogen is abundant,
CH3CHO serves as the sink for carbon chemistry.
In
the outer region where the temperature is low enough
for C2H3 to stay on the grain surface and hydrogenate
to C2H5, we are seeing CH3OH and C2H5 as the end
products of COM chemistry.
The reaction pathways described above demonstrate
how H2CCO, CH3CHO, C2H2 and CH3OH stand in for
complex organic molecules in our model.
In reality,
organic molecules may grow more complex as ketene,
acetaldehyde, ethyl, and methanol ice react with other
species in ways not included in our model. Despite our
upper limit of two carbon atoms per molecule, the net
movement of carbon from CO to ices should be a robust
result for ionized regions that contain CH4 gas. In or-
der to demonstrate that our reaction network does not
falsely predict carbon depletion from the gas phase, we
investigate an alternative network with a low ketene for-
mation rate in the next section. We also compare our
model abundances with results from the c2d (cores to
disks) Spitzer Legacy Program as summarized by Oberg
et al. (2011a), and discuss the effect of different assump-
tions about grain evolution.
4.5. Model dependence on reaction rates, initial
conditions, and grain evolution models
Our model results depend on the disk physical condi-
tions, initial cloud phase abundances, and reaction rates.
In this section, we present a detailed study of how results
depend on reaction rates, initial abundances, and grain
evolution.
In section 4.5.1, we verify that our ketene
ice sink does not artificially remove carbon from the gas
phase by running a model with the ketene formation
rate decreased by 10 orders of magnitude.
In Section
4.5.2, we compare the abundances from our molecular
cloud preprocessing model with the ice abundances ob-
served by the Spitzer c2d team in the envelopes of low-
mass protostars ( Oberg et al. 2011a). In Section 4.5.3,
we investigate the effects of grain evolution on the disk
temperature and chemistry.
Our current model cannot consider gaps or inner holes
without significant modifications. The effects of such
structures on chemical evolution requires further study.
4.5.1. Models with low H2CCO formation rates
The formation of carbon sinks such as ketene greatly
reduces the gas-phase CO abundance in our model and
suggests that observations of CO depletion may trace
complex-molecule formation instead of freezeout of CO.
In this section, we test the robustness of our result by
artificially suppressing ketene formation rates. The CO
abundance is independent of the exact form of carbon-
bearing ice as long as our model does not artificially
remove carbon from the gas phase. However, having
H2CCO ice instead of C2Hx ice as a sink in the model
can potentially reduce the amount of oxygen in the gas
phase, therefore affecting the CO abundance. To test
the effect of H2CCO formation rate on CO abundance,
we ran the chemical model with H2CCO formation rates
artificially turned down by ten orders of magnitude while
keeping other parameters the same.
We compare output abundances at the end of the 3
Myr evolution of this experiment with those from our
fiducial COM-forming model (described in Section 4.3)
in Table 5.
The H2CCO formation is strongly suppressed as ex-
pected, resulting in a negligible abundance. Without the
pathways that convert hydrocarbons (C2Hx) to ketene,
the abundance of hydrocarbons (C2Hx) is significantly
higher, but most other abundances show little change.
The gas-phase CO abundance increases from 13.6% to
17.3% of the elemental carbon.
To summarize, the choice of end-member species in
the chemical reaction network can affect the gaseous CO
abundance. Translating observed CO isotopologue line
intensities into disk surface densities would include an
uncertainty of at least a factor of two for any given disk
annulus. However, the effect of the ketene sink on the
CO gas abundance in our modeled 70 AU disk as a whole
is small. One would be able to translate a measurement
of the mass of C17O or C18O gas in the planet-forming
region of a disk into a total mass available for giant
planet formation without large uncertainties due to the
chemistry of complex organic molecule formation.
4.5.2. Comparing with the c2d cloud abundances
We test the sensitivity to initial conditions in our
chemical reaction network by comparing the abundances
at the end of our molecular cloud preprocessing model
with measured ice abundances in low-mass protostel-
lar envelopes. Oberg et al. (2011a) combined Spitzer/
IRS spectra for about 50 low-mass protostars with in-
frared ice features with near-infrared ground-based ob-
servations of ice features such as 3µm H2O, 4.65µm CO,
and 3.53µm CH3OH, presenting an overview of the ice
inventory during the embedded phase of star formation.
Because the absolute ice abundances vary from source
to source, Oberg et al. (2011a) presented the median ice
abundance ratio with respect to water ice abundance in
the combined sample. The chemical compositions ob-
served in low-mass protostellar envelopes are expected
to be similar to those at the end of the molecular cloud
phase, because the materials are not yet heavily heated
by the central star or processed by shocks.
We compare c2d observed ice abundance ratios and
outputs of our cloud phase model (as described in Sec-
tion 3.1) in the first two columns of Table 4. The CO2,
CH4, and CO abundances with respect to the water ice
10
abundance in our cloud model are about 1.2, 5, and 0.5
times the values observed in nearby star-forming regions,
respectively ( Oberg et al. 2011a). While not an exact
match, our computed abundances relative to water ice
are of the same order of magnitude as observed values
for major carbon-bearing species.
To test how our disk model varies based on small
changes in input abundances, we conduct an experiment
using the Oberg et al. (2011a) observed abundances as
the initial conditions for the chemical evolution, while
keeping other parameters the same. We keep the same
abundances as predicted by our molecular cloud model
for H2O ice and minor species not observed by the c2d
team, but scale the abundances of other c2d-observed
ices to match the c2d results. The abundances pre-
dicted by the molecular cloud model, and the abun-
dances adjusted according to c2d ratios, are listed in the
last two columns of Table 4. This experiment is only
to demonstrate the model dependence on initial input
abundances. We do not conserve the total number of C
and O atoms per hydrogen atom by artificially scaling
the molecular abundances.
The change of input abundances does not change CO
abundance significantly in our experiment. Integrated
over the entire disk, the total available carbon in the
form of CO gas increases from 13.6% to 18.3%, while
the carbon locked in COM decreases from 44.8% to
37.0% However, the dominant kind of COM is different
from that of the standard model. Due to the signifi-
cantly lower input methane abundance, the percentage
of molecules that contain two carbon atoms is only one
third of the value in the fiducial model, and the methanol
abundance is three times the original value. The paucity
of ices that contain two carbon atoms when the initial
methane abundance is low suggests that even though
the formation of the carbon-carbon bond relies on both
CH4 and C+ liberated from CO, the formation of car-
bon chain depends largely on the methane abundance
the disk inherits from the cloud.
We find that the abundances predicted by our molec-
ular cloud model agree with observational results within
a factor of two. In our experiment, adjusting the abun-
dances inherited by the disk to match observed abun-
dances in cold protostar envelopes can lead to a higher
abundance of methanol and a smaller overall COM
abundance than in our fiducial model. However, the
computed CO abundance is only 1.35 times higher.
4.5.3. Different models of grain evolution
In the fiducial model, we assume 90% by mass of the
of dust has already grown to larger than mm size, and
use a gas to dust ratio of 1000 throughout. However,
grain evolution could be slower than what we assume.
Specifically, the disk could have a larger portion of dust
than what we assumed at the beginning of the evolu-
tion. Small dust grains provide most of the opacity
that is important for heating and attenuating UV and x-
ray radiation, and they also provide most of the surface
area available for freeze-out and grain surface reactions.
Miotello et al. (2014) have found that the evolution of
dust grains can change the disk opacity, which can fur-
ther increase the significance of isotopologue-selective
photodissociation, and change the CO fractionation.
To investigate the effect of grain evolution, we con-
struct a series of models with gas to dust ratio gradually
evolving from 100 to 1000, assuming the same opacity
profile for the dust grains. We choose to change the gas
to dust ratio to trace the loss of solid material to bigger
bodies, rather than changing the grain size distribution,
because a balance between collisional aggregation and
fragmentation yields a size distribution whose slope at
the small end is insensitive to the dust abundance (Birn-
stiel et al. 2011).
In the models considering grain evolution, the total
dust opacities are larger than those in the fiducial model,
resulting in lower disk temperature. On average, the
disk temperature is 10% lower at 100 yr where the gas-
to-dust ratio is 10 times the fiducial value, and 5% lower
at 1 Myr when the gas-to-dust ratio is one third of the
fiducial value.
The output abundances averaged over the entire disk
are presented in the last column of Table 4. Despite the
decrease of temperature, the carbon content in the grain
evolution model is very similar to that in our fiducial
model by the end of the 3 Myr evolution. This is be-
cause CO depletion happens on a Myr time scale when
grain evolution has already started, and the grain sur-
face area is not limiting the freeze-out time scale as long
as some dust grains are present. The final CO abun-
dance relative to total carbon decreases from 13.6% in
the standard model to 11.8% in this model, a change in
the opposite direction to that found in the other model
variations. The difference between the two models is
largest at 1 Myr, when the CO/C ratio is 27.5% in the
standard model and 21.8% in the grain evolution model.
Unlike in Miotello et al. (2014), considering grain evo-
lution does not significantly change abundance ratios of
CO isotopologues in our models. We defer detailed dis-
cussions on isotopic ratios to section 4.6.
4.6. Fractionation of CO isotopologues
In addition to understanding the formation and de-
struction pathways of CO, one also needs to know how
C and O isotopes fractionate among different molecules
in order to link 13CO, C17O, and C18O observations with
disk masses. The isotopologue ratio in CO and the ma-
jor carbon sink H2CCO are presented in Figure 9. The
CO/13CO ratio is close to the elemental isotopic value
11
of 77.15 in the inner 15 AU where CO gas is abundant.
In regions where CO is depleted, the CO/13CO ratio
is about 45, much smaller than the input atomic abun-
dance ratio. The low CO/13CO ratio is inherited from
the molecular cloud model.
In the inner disk, where new CO molecules are form-
ing from recycled ketene that originally gained its car-
bon atoms from both CH4 and CO, the CO/13CO ratio
approaches the input, atomic C/13C ratio. Integrated
over the entire disk, the average CO/13CO ratio is 59.15.
Disk mass estimates based on optically thin 13CO rota-
tional emission (possible beyond ∼ 18 AU; see Section
5) should account for isotopic fractionation.
In the shielded interior of our disk, we predict
C16O/C17O and C16O/C18O ratios similar to those
measured by Smith et al. (2009) from observations of
the envelope of protostar Reipurth 50. Our computed
CO/C17O and CO/C18O ratios are slightly smaller than
the input 16O/17O = 2300 and 16O/18O = 500 ratios
throughout most of the disk, though the recycling of
ketene back to CO pushes C16O/C17O and C16O/C18O
higher near the disk surface. The CO oxygen isotope
fractionation identified in our model has an average ef-
fect of less than 14% (averaged over mass). We find that
oxygen isotope fractionation would not significantly bias
disk mass estimates based on C17O or C18O emission.
We also caution against over-interpretation of the oxy-
gen isotope fractionation found in this model, as vari-
ations in UV flux that are well within observed ranges
(France et al. 2014) can affect isotope ratios. In Miotello
et al. (2014), UV radiation is not attenuated efficiently
in disks with only larger grains. Rare CO isotopologues
only exist in a smaller region further below the disk sur-
face due to photo-dissociation, resulting in larger CO to
rare CO ratios. In contrast, our models include a pop-
ulation of small dust grains even for models with grain
evolution (Section 4.5.3), and therefore the UV radia-
tion is sufficiently absorbed. Moreover, following Landry
et al. (2013), our disk models only follow the disk up
to about two scale heights, and do not follow the disk
atmosphere, where photochemical reactions matter the
most. This explains why unlike in Miotello et al. (2014),
considering grain evolution does not significantly change
the CO isotopologue ratios.
Having verified that our predicted 13CO, C18O, and
C17O abundances are robust results that (1) do not de-
pend heavily on our reaction network choice and (2)
closely track observed values, we are now ready to cal-
culate the optical depths of rotational emission lines.
5. OPTICAL DEPTH OF ROTATIONAL EMISSION
FROM CO ISOTOPOLOGUES
Our goal is to demonstrate that rare CO isotopologues
can provide a window into the planet formation zone
12
through their rotational emission lines. The ideal emis-
sion line would be optically thin all the way to the disk
midplane, but still detectable with ALMA. We calcu-
late the optical depth of CO rotational emission lines at
each disk radius by integrating the absorption efficiency
through the disk. We first calculate the total absorption
efficiency across the emission line as
(cid:90)
kνdν =
hνij
c
(niBij − njBji) ,
in which
Bji =
Bij =
Bij,
gi
gj
8π3
3h2 µij2,
ni = finx.
(32)
(33)
(34)
(35)
(cid:90)
kν is the absorption coefficient, ni, and nj are the num-
ber of molecules in states i and j (i is the lower state),
and fi and fj are the fraction of molecules in states i and
j. gi and gj are the degeneracies; νij is the frequency of
the transition; µij is the matrix element of the electric
dipole moment for the transition; and Bij and Bji are
the Einstein coefficients of the transition. Substituting
into equation 32, we have:
(cid:20)
(cid:18)−hνij
(cid:19)(cid:21)
.
kT
kνdν =
hνij
c
finx
8π3
3h2 µij2
1 − exp
(36)
We take fi as the local thermodynamic equilibrium
(LTE) value set by the Boltzmann distribution.
To estimate the absorption at the line center k0, we
divide the total absorption efficiency by the full-width
half-maximum ∆ν of the Doppler broadening profile:
∆ν = 2.355 ×
ν2
0 kT
M c2 ,
(37)
where M is the mass of the molecule. We then have
(cid:114)
(cid:82) kνdν
∆ν
k0 =
.
(38)
Finally, we compute face-on optical depths as a function
of distance from the star by integrating the absorption
efficiency k0 over the disk's vertical direction as
(cid:90) zsurf ace
0
τ0 =
k0dz.
(39)
The optical depths of J= 1 → 0 lines for CO isotopo-
logues are shown in the upper panels of Figure 10. The
upper-left panel shows optical depths after 100 years of
disk evolution (i.e., a disk surrounding a just-emerged
T-Tauri star), and the upper-right panel shows the disk
after 3 Myr of evolution. We highlight the J= 1 → 0
transition because the temperature of CO gas is rela-
tively high compared to the energy needed to excite the
J = 1 rotational level (5.5K), therefore J = 1 → 0 line is
the most optically thin line among low order rotational
lines observable with ALMA. We can immediately see
that C17O traces the disk midplane outside 8 AU, and
C18O traces the midplane outside 12 AU. Observations
of C17O and C18O emission from nearby disks would
provide estimates of the amount of mass available to
form planets like Uranus and Neptune, and place a lower
limit on the mass available to form Jupiter and Saturn.
The optical depth of C17O as a function of radius is
plotted for several epochs in the lower left panel of Fig-
ure 10. The optical depth is a strong function of radius
and time. The optical depth in the inner part of the
disk increases as the disk evolves due to the dissocia-
tion of CO2 into CO, while the optical depth beyond
15 AU first increases, but then decreases and falls below
the initial value due to the formation of more complex
carbon-bearing molecules (sections 4.2, 4.3). At 3 Myr
the optical depth remains above one out to about 42 AU
for CO, 18 AU for 13CO, 12 AU for C18O, and 8 AU for
C17O. The possibility that CO abundance changes over
time due to chemical reactions means that disk masses
measured by CO emission are degenerate with age.
Continuum emission from dust also contributes flux
at the frequencies of CO emission lines. Even if the
emission from a CO isotopologue is optically thin, high
optical depth in the dust could prevent us from see-
ing the disk midplane. To estimate the optical depth
contributed by the dust, we integrate the dust opacity
at the relevant wavelength over the disk height. We
use opacities from Semenov et al. (2003), for grains
with a 5-layered sphere topology. As in the thermo-
dynamic model (see Section 2 for details), we use a gas
to dust mass ratio of 1000. We assume the dust has
undergone some degree of grain growth, with remain-
ing solids locked into larger bodies, consistent with our
RADMC dust radiative transfer models. Optical depth
contributed by dust is plotted in the lower right panel of
Figure 10. The optical depths contributed by dust are
almost the same for all isotopologues due to the small
shift in frequency between emission lines, so we are only
showing the frequency at the center of CO transitions.
The optical depths contributed by dust are smallest for
the J= 1 → 0 line due to lower frequency - less than
one beyond 2 AU, and less than 0.1 beyond 10 AU. The
dust optical depth for higher order transitions become
less then one within 10 AU.
For low-mass disks that have experienced some grain
growth, C17O is a promising tracer of the disk midplane
in the giant planet-forming region. Outside of 10 AU,
where analogs to the Kuiper Belt may be forming (Bry-
den et al. 2009), C18O and 13CO may be useful midplane
tracers. Note, however, that our fiducial model is a low-
mass disk (0.015 M(cid:12)), and the actual midplane locations
traced by CO gas may therefore fall somewhat outside
of 10 AU. However, our results demonstrate the value
of observing CO isotopologues to reconstruct the mass
distributions in the inner 30 AU of nearby disks.
6. CONCLUSIONS
Our chemical model of an MRI-active protoplanetary
disk has led to the following conclusions.
CO does not freeze out anywhere in our modeled re-
gion - the inner 70 AU of the disk - due to efficient heat-
ing by stellar irradiation. Instead, the abundance of CO
is a complex function of both radius and time. This de-
pendence must be modeled correctly in order to deduce
gas properties from observations of CO isotopologues.
The fate of CO is tied up in the formation of complex
organic molecules. While the detailed chemical evolu-
tion depends on the input abundances from the molecu-
lar cloud and which species are included in the network,
the main result is robust to these variations. Different
models of grain evolution also produce small changes
in the outcomes. For three different chemical assump-
tions and two different scenarios for grain evolution, the
partition of carbon, integrated over the modeled inner
70 AU of the disk, ranges from 11.8% to 18.3% in CO
gas, 31.0% to 38.4% in CO2 ice, and 37.0% to 45.0% in
complex organic molecules.
Fractionation of oxygen isotopes appears not to play
a major role in C17O abundances, though stars with un-
usually strong UV accretion luminosity may have disks
depleted in C17O. The optical depth of low-J rotational
lines of C17O are around unity in the giant planet form-
ing region, while more common isotopologues are quite
opaque. With our computed C17O/H2 abundance as
a function of radius and an age, one would be able to
translate the observed C17O line intensity into available
planet-forming mass, to within a factor of a few. The
optical depth of dust in the CO isotopologue (J= 1 → 0)
transition wavelengths is small enough to allow observa-
tions of the mid-plane beyond a few AU; for higher tran-
sitions the dust will be opaque within 5-8 AU. However,
this result relies on our assumption of grain evolution, in
which 90% of the solid mass has accumulated in pebbles
and larger solid objects at the start of disk evolution at
0.1 Myr, but our experiment with slower evolution of
the gas to dust ratio produce similar results in the end.
Looking ahead to future work, we emphasize that
emission lines from the CO isotopologues all have dif-
ferent optical depths and can probe different vertical
layers of a target disk. Comparing line profiles of
emission from multiple isotopologues could reveal ver-
tical variations in turbulent velocity, and therefore con-
strain the angular momentum transfer mechanism that
drives the disk evolution. The current leading angu-
lar momentum transfer mechanism -- MRI turbulence --
predicts non-turbulent "dead" zones on the disk mid-
13
plane, which leads to a decrease of the turbulent ve-
locity toward the midplane (Fromang & Nelson (2006),
Simon et al. (2011)). However, new models by Bai
& Stone (2013) and Gressel et al. (2015) suggest that
magnetocentrifugal winds are more likely to drive ac-
cretion than the MRI. Aside from possibly driving ac-
cretion, turbulence also determines the behavior of dust
grains to a great extent. An observational investiga-
tion of gas velocities in the midplanes of planet-forming
disks would have profound implications for planetesimal
growth models. ALMA is able to detect emission from
multiple CO isotopologues in one observation. With its
new long baselines and large collecting area, it is sen-
sitive to the inner regions of disks, which have small
emitting areas but contribute appreciably to line wings.
CO observations may be able to probe not only the mass
distribution in planet-forming regions, but the gas dy-
namics as well.
We conclude that CO isotopologue rotational emission
can probe the mass distribution and chemical evolution
of the inner radii of protostellar disks, allowing observers
to peer directly into planet nurseries.
Acknowledgments: Work by MY, KW, SDR and NJT
was supported by NASA grant NNX10AH28G. NJE and
MY were supported in part by NSF Grant AST-1109116
to the University of Texas at Austin. This work was
performed in part at the Jet Propulsion Laboratory,
California Institute of Technology. NJT was supported
by grant 13-OSS13-0114 from the NASA Origins of So-
lar Systems program. We are grateful to the referee
for helpful suggestions. We would like to thank Ed-
win Bergin, Jacob Simon, Ilse Cleeves, Jeong-Eun Lee,
Seok-Ho Lee, Jeffrey Cuzzi, Paul Estrada, Colette Salyk,
Karin Oberg, and Raquel Salmeron for useful discus-
sions.
Alexander, R. D., Clarke, C. J., & Pringle, J. E. 2006, MNRAS,
Gressel, O., Turner, N. J., Nelson, R. P., & McNally, C. P. 2015,
369, 229
ApJ, 801, 84
REFERENCES
Ali-Dib, M., Mousis, O., Petit, J.-M., & Lunine, J. I. 2014a, ApJ,
785, 125
-- . 2014b, ApJ, 793, 9
ALMA Partnership, Brogan, C. L., P´erez, L. M., et al. 2015,
ApJL, 808, L3
Bai, X.-N., & Goodman, J. 2009, ApJ, 701, 737
Bai, X.-N., & Stone, J. M. 2013, ApJ, 769, 76
Balbus, S. A., & Hawley, J. F. 1991, ApJ, 376, 214
Banzatti, A., Testi, L., Isella, A., et al. 2011, A&A, 525, A12
Bergin, E. A., Cleeves, L. I., Gorti, U., et al. 2013, Nature, 493,
644
Birnstiel, T., Ormel, C. W., & Dullemond, C. P. 2011, A&A,
525, A11
Bryden, G., Beichman, C. A., Carpenter, J. M., et al. 2009, ApJ,
705, 1226
Chiang, E. I., & Goldreich, P. 1997, ApJ, 490, 368
Cleeves, L. I., Adams, F. C., & Bergin, E. A. 2013a, ApJ, 772, 5
Cleeves, L. I., Adams, F. C., Bergin, E. A., & Visser, R. 2013b,
ApJ, 777, 28
D'Antona, F., & Mazzitelli, I. 1994, ApJS, 90, 467
Dodson-Robinson, S. E., & Bodenheimer, P. 2010, Icarus, 207,
491
Dodson-Robinson, S. E., Willacy, K., Bodenheimer, P., Turner,
N. J., & Beichman, C. A. 2009, Icarus, 200, 672
Dullemond, C. P., & Dominik, C. 2004, A&A, 421, 1075
Dunham, M. M., & Vorobyov, E. I. 2012, ApJ, 747, 52
France, K., Schindhelm, E., Bergin, E. A., Roueff, E., & Abgrall,
H. 2014, ApJ, 784, 127
Fromang, S., & Nelson, R. P. 2006, A&A, 457, 343
Garaud, P., Meru, F., Galvagni, M., & Olczak, C. 2013, ApJ,
764, 146
Garmire, G., Feigelson, E. D., Broos, P., et al. 2000, AJ, 120,
1426
Garrod, R. T., & Pauly, T. 2011, ApJ, 735, 15
Graedel, T. E., Langer, W. D., & Frerking, M. A. 1982, ApJS,
48, 321
Grevesse, N., & Sauval, A. J. 1998, SSRv, 85, 161
Guilloteau, S., Dutrey, A., Pi´etu, V., & Boehler, Y. 2011, A&A,
529, A105
Habing, H. J. 1968, BAN, 19, 421
Henning, T., & Stognienko, R. 1996, A&A, 311, 291
Hollenbach, D., Kaufman, M. J., Bergin, E. A., & Melnick, G. J.
2009, ApJ, 690, 1497
Igea, J., & Glassgold, A. E. 1999, ApJ, 518, 848
Isella, A., Carpenter, J. M., & Sargent, A. I. 2010, ApJ, 714, 1746
Kraus, A. L., & Ireland, M. J. 2012, ApJ, 745, 5
Landry, R., Dodson-Robinson, S. E., Turner, N. J., & Abram, G.
2013, ApJ, 771, 80
Langer, W. D., & Penzias, A. A. 1993, ApJ, 408, 539
Lee, H.-H., Herbst, E., Pineau des Forets, G., Roueff, E., & Le
Bourlot, J. 1996, A&A, 311, 690
Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ, 217,
425
Miotello, A., Bruderer, S., & van Dishoeck, E. F. 2014, A&A,
572, A96
Oberg, K. I., Boogert, A. C. A., Pontoppidan, K. M., et al.
2011a, ApJ, 740, 109
Oberg, K. I., Murray-Clay, R., & Bergin, E. A. 2011b, ApJL,
743, L16
Oliveira, I., Pontoppidan, K. M., Mer´ın, B., et al. 2010, ApJ,
714, 778
P´erez, L. M., Carpenter, J. M., Chandler, C. J., et al. 2012,
ApJL, 760, L17
P´erez, L. M., Chandler, C. J., Isella, A., et al. 2015, ApJ, 813, 41
Piso, A.-M. A., Oberg, K. I., Birnstiel, T., & Murray-Clay, R. A.
2015, ApJ, 815, 109
Pollack, J. B., McKay, C. P., & Christofferson, B. M. 1985,
Icarus, 64, 471
Qi, C., D'Alessio, P., Oberg, K. I., et al. 2011, ApJ, 740, 84
Qi, C., Oberg, K. I., & Wilner, D. J. 2013, ApJ, 765, 34
Robrade, J., Gudel, M., Gunther, H. M., & Schmitt, J. H. M. M.
2014, A&A, 561, A124
Semenov, D., Henning, T., Helling, C., Ilgner, M., & Sedlmayr,
Gredel, R., Lepp, S., Dalgarno, A., & Herbst, E. 1989, ApJ, 347,
E. 2003, A&A, 410, 611
289
Shakura, N. I., & Sunyaev, R. A. 1973, A&A, 24, 337
14
Simon, J. B., Armitage, P. J., & Beckwith, K. 2011, ApJ, 743, 17
Smith, R. L., Pontoppidan, K. M., Young, E. D., Morris, M. R.,
& van Dishoeck, E. F. 2009, ApJ, 701, 163
Tazzari, M., Testi, L., Ercolano, B., et al. 2015, ArXiv e-prints,
Umebayashi, T., & Nakano, T. 1981, PASJ, 33, 617
-- . 2009, ApJ, 690, 69
van Dishoeck, E. F., & Black, J. H. 1988, ApJ, 334, 771
Visser, R., van Dishoeck, E. F., & Black, J. H. 2009, A&A, 503,
arXiv:1512.05679
323
Telleschi, A., Gudel, M., Briggs, K. R., Audard, M., & Scelsi, L.
2007, A&A, 468, 443
Williams, J. P., & Cieza, L. A. 2011, ARA&A, 49, 67
Woodall, J., Ag´undez, M., Markwick-Kemper, A. J., & Millar,
Thommes, E. W., Duncan, M. J., & Levison, H. F. 2002, AJ,
T. J. 2007, A&A, 466, 1197
123, 2862
Woods, P. M., & Willacy, K. 2009, ApJ, 693, 1360
Tsiganis, K., Gomes, R., Morbidelli, A., & Levison, H. F. 2005,
Nature, 435, 459
15
Table 1. A summary of parameters and definitions of variables
Symbol
ISRF
COM
MRI
Stellar mass
Disk mass
nH2
fH
Abundance
A(B)
T
Teq
Tacc
ζabs
ζsca
LX,29
Σcr
ζcr
-
Definition or Value
Interstellar radiation field
1ISRF = 1.6 × 10−3ergs−1cm−2 for the ionizing UV radiation
Complex organic molecules
Magneto-rotational instability
0.95 M(cid:12), we follow the stellar evolution from 0.1 Myr (T(cid:63) = 4600K, and R(cid:63) = 5.5 R(cid:12)),
to 3 Myr (T(cid:63) = 4500K, and R(cid:63) = 1.5 R(cid:12)).
0.015 M(cid:12) within 70 AU from the central star
number density of hydrogen molecules
0.735, mass fraction of hydrogen
number density with respect to the number density of hydrogen nuclei (nH + 2nH2 ).
A(B) stands for A×10B
disk temperature (same for dust and gas)
equilibrium temperature contributed by stellar irradiation
accretion temperature contributed by the MRI turbulence
6 × 10−12 s−1, ionization rate for direct x-ray absorption
10−15 s−1, ionization rate for scattered x-ray photons
20, stellar X-ray luminosity in units of 1029 ergs−1
96 gcm−2, characteristic column density of attenuation for cosmic rays
cosmic ray ionization rate, 1.2 × 10−17 s−1 for H2 molecules when calculating the
reactions lead by cosmic ray induced photons, the total ionization rate is taken as
1.3 × 10−17 s−1.
Table 2. Input abundances of the molecular cloud model
Species Abundance 1,2
H
C+
O
17O
Si+
1.00 (-2)
7.21 (-5)
1.76 (-4)
7.64 (-8)
3.00 (-9)
Element Abundance
He
13C+
18O
N
1.40 (-1)
9.34 (-7)
3.51 (-7)
2.14 (-5)
1 Number density with respect to the number density of hydrogen nuclei (nH + 2nH2 ). The abundances are taken from Graedel
et al. (1982).
2 A(B) stands for A×10B
16
Table 3. Binding energies and ice locations for carbon-bearing molecules
Species
H2O
CO2
CO
CH4
Species
CH3
CH3OH
C2H2
C2H3
C2H4
C2H5
C2H6
H2CCO
CH3CHO
H2CO
Binding Energy (K) 1,2
Ice line (100 yr) 3
Ice line (3 Myr) 3
5.77 (3)
2.86 (3)
8.55 (2)
1.08 (3)
1.5 AU
18 AU
> 70 AU
> 70 AU
1 AU
2 AU
> 70 AU
> 70 AU
Binding Energy (K)
Ice location (3 Myr) 4
1.16 (3)
4.23 (3)
2.40 (3)
1.76 (3)
2.01 (3)
2.11 (3)
2.32 (3)
2.52 (3)
2.87 (3)
1.76 (3)
r= 5 − 20AU, z=0.2 − 2AU, and r> 40AU
2 − 25 AU surface
> 40AU
most of the disk 5 − 45AU
r= 8 − 25AU, z= 0.5 − 2.5AU
1 Binding energy is cited in energy (E in erg) divided by Boltzmann's constant (kB in erg/K) for easy interpretation, therefore
has a unit of Kelvin.
2 A(B) stands for A×10B
3 The ice line (condensation front) location on the disk midplane.
4 For CH3OH, C2H2, C2H5, H2CCO, and CH3CHO, we mark the location where we see certain ice species, which de-
pend on both the condensation temperature and the ability to form the molecule. The change of abundance can be
very gradual. Here we quote the boundary location where 10% of carbon is stored in corresponding species. We are only
showing results for the end of evolution (3 Myr) because those molecules are not formed until about 1 Myr into the disk evolution.
Table 4. Abundances at the end of the cloud phase
c2d fraction 2 MC abundance 3
Ice species MC fraction 1
H2O
CO2
CH4
CO
Gas species
CO
100
35
24
15
100
28
5
29
NA 5
9.1(−5)
3.2(−5)
2.2(−5)
1.3(−5)
c2d abundance 4
9.1(−5)
2.7(−5)
4.6(−6)
2.7(−5)
1.9(−6)
1.9(−6)
1 Fractional abundances in our molecular cloud model, normalized to the H2O ice abundance.
2 Fractional abundances in the c2d result, normalized to the H2O ice abundance.
3 Output abundances in our molecular cloud model.
4 Abundances scaled according to the c2d result, using the H2O abundance as the reference.
5 Oberg et al. (2011a) does not include gaseous species.
17
Table 5. Output abundances from different models
Ice
CO2
CH3OH
CH3CHO
H2CCO
4
C2Hx
COM 5
Gas
CO
CH4
Standard1 Low H2CCO2
36.2
8.1
3.9
15.2
16.1
44.8
13.6
1.6
37.6
7.4
4.4
0.0
25.9
39.4
17.3
1.7
c2d3
31.0
24.6
5.6
3.4
3.1
37.0
18.3
0.2
evolving D/G
38.4
10.1
3.01
9.0
21.0
45.0
11.8
1.9
1 Model with original reaction rates and the input from our cloud model abundances. The percentage of total available carbon
contained in each species is shown in the table. All the results are values integrated over the entire disk at the end of our
modeled disk evolution at 3 Myr.
2 Model with small H2CCO formation rates and the input from our cloud model abundances.
3 Model with original reaction rates and the input adjusted to the c2d fractional abundances. See Table 4 for details.
4 Hydrocarbons that contain two carbon atoms.
5 The total abundance of complex organic molecules, including CH4, CH3OH, H2CCO, CH3CHO, and C2Hx.
18
Figure 1. Workflow of our disk model
Blue boxes show three major components of our disk model. Green boxes show the model inputs taken directly from the
literature, and yellow boxes are showing intermediate results computed in our study. Quantities in boxes with black borders
evolve as a function of time, and the quantities in boxes without borders do not change over time.
The arrows and variables next to the arrows are showing how information is passed between different model components.
19
Reaction Rate: UMISTInput: MC abundances UV, Xrays, Cosmic RaysDensity TempAtomic AbundanceMolecular AbundanceLine Optical DepthsChemical Model1+1D MRIThermodynamical ModelRADMC2D Opacity𝞺𝞺𝞺, TaccTeqStellar Evolution (R*, T*)Figure 2. Opacity as a function of wavelength for the 5-layer multishell grain model of Semenov et al. (2003). The grain
temperature used for this plot is 155 K.
20
Figure 3. Upper left panel: Disk temperature (100yr) as a function of the radial distance to the central star (R) and height above
the disk midplane (Z); Upper right panel: Disk temperature (3Myr); Lower left panel: Accretion heating (Tacc) contributed by
the MRI turbulence (3Myr); The patches on the three color coded plots are interpolation artifacts, although the "arc" of light
blue at ∼ 1AU above the midplane going from 5 − 30 AU is real heating due to the MRI turbulence. Lower right panel: A
comparison of the midplane temperature (Teq only) computed by RADMC and the 1+1D model with Rosseland mean opacities
(3Myr). The 1 + 1D model does not capture the heating contributed by longer wavelength radiation and underestimates the
disk temperature in the inner 20AU. This effect is more severe at smaller radius due to larger optical depth, resulting in a lower
temperature at smaller radius on the disk midplane - even though its surface temperature is higher.
21
010203040506070R (AU)0246810121416Z (AU)Accretion Heating (K) (3 Myr)0.00.51.01.52.02.53.03.54.00246810R (AU)0.00.51.0Z (AU)010203040506070R (AU)020406080100120140Z (AU)Midplane Temperature (K)(3 Myr)1+1D Rosseland MeanRADMCFigure 4. Estimated ionization rates contributed by various mechanisms at the end of our disk evolution. Upper left panel:
UV ionization rate (an order of magnitude estimation for the H2O molecule, the ionization rate in the disk interior is lower
than 10−20s−1, and therefore does not show up in the plot.); Upper right panel: X-ray ionization; Lower left panel: cosmic
ray ionization (an order of magnitude estimation considering only the H2 molecule); Lower right panel: fraction contributed by
X-ray ionization. X-ray ionization dominates at z/r > 0.1, and cosmic rays account for most of the ionization for z/r < 0.1,
where the UV and X-ray radiation from the central star are sufficiently attenuated.
22
Figure 5. Fractional abundances at the end of the 3Myr evolution (the number density with respect to the number density of
hydrogen nuclei, nH + 2nH2 ). The color scale is in logarithm. Upper left: CO exists in a large abundance for r< 15AU; Upper
right: CO2 ice exists in most part of the disk where the temperature is low enough for it to stay on grain surfaces; Lower left:
H2CCO (ketene) ice is the major carbon sink beyond 15 AU from the central star; Lower right: Other complicated organic
molecules such as C2Hx, and CH3CHO (acetaldehyde) exist in a layer between 10 − 30 AU, closer to the disk surface.
23
Figure 6. Fractional abundances continued. The figure setup is the same as in Fig. 5. Upper left: C2H2 is abundant where
the temperature is too high for any carbon-bearing ices to freeze out; Upper right and lower left: CH3OH and CH3CHO serve
as the carbon sinks where the temperature is high enough to evaporate H2CCO (ketene) ice; Lower right: C2H5 is able to form
where the temperature is low enough for C2H3 to stay on grain surface and hydrogenate.
24
Figure 7. Abundances of major carbon-bearing molecules as functions of time. Upper panel: 38 AU on the disk midplane; lower
panel: 60 AU on the disk midplane. Points show the values of actual data points in our models.
25
0.00.51.01.52.02.53.00.00.51.01.52.0Time (Myr)Abundance(10−5)Time evolution of abundances (40AU, midplane)COCH4C2H2 iceH2CCO iceC2H5 iceCH3OH ice0.00.51.01.52.02.53.00.00.51.01.52.0Time (Myr)Abundance(10−5)Time evolution of abundances (60AU, midplane)COCH4C2H2 iceH2CCO iceC2H5 iceCH3OH iceFigure 8. Reaction network for major carbon-bearing species. The species with boxes drawn around them are sinks, and letter
G denotes species that are frozen out on grain surfaces. Starting from the upper left side the the figure: (1) The dissociation of
CO is initiated by He+; (2) The topmost pathway shows methanol formation in relatively hot regions of the disk, where H2CCO
molecules are not able to stay on the the grain surface. (3) The blue lines represent processes that move carbon from C2H3
to GCH3CHO (acetaldehyde); (4) The green lines trace a path for removing carbon from ketene in warm parts of the disk; To
summarize, C2H2 gas exists where the temperature is too high for icy carbon sinks to form. GCH3CHO and GCH3OH are the
carbon sink in warm regions, GH2CCO in the majority part of the disk where the temperature is lower.
26
COC+C2H+He+HeOCH3CH4CH4C2H3+C2H2+HCO+C2H5+C2HC2H2e-e-freezeGC2HfreezeGC2H2GHGHGC2H3C2H3desorbfastGHGC2H4slowGHGC2H5CH3H2COfreezeGH2COGHGHGCH2OHGCH3OHH2CCOGH2CCOfreezeC+CH2CO+CH2COe-C2C2H5CH3CHOdesorbfreezeGCH3CHOOOC2H4+e-Figure 9. Upper left: CO/13CO ratio (all four plots are for the end of the 3 Myr evolution); Upper right: C/13C ratio in
H2CCO ice; Lower left: CO/C17O ratio; Lower right: CO/C18O ratio. Unsuperscripted species denote the most common
isotopologue.
27
Figure 10. Upper left panel: Optical depth of various CO isotopologues J = 1 → 0 (100 yr). Upper right panel: Optical
depth of various CO isotopologues J = 1 → 0 (3 Myr). Lower left panel: Optical depth of C17O J = 1 → 0 (time evolution).
The optical depth increases over time roughly in r < 20 AU due to the photodissociation of CO2 lead by cosmic ray-induced
photons, and the optical depth at r > 20 AU decreases over time due to the formation of COMs. Lower right panel: Optical
depth contributed by dust emission at the wavelength of various CO transitions. Dust emission should not contribute much to
observed fluxes in low-order CO rotational emission lines beyond 10 AU.
28
Optical depth of CO J = 1-0 (100 yr)0102030R (AU)-4-2024Optical Depth (log10)CO13COC18OC17O13C18OOptical depth (J=1-0, 3Myr)0102030R (AU)-4-2024Optical Depth (log10)CO13COC18OC17O13C18OOptical depth of C17O (J=1-0)0102030R (AU)-3-2-101Optical Depth (log10)100 yr1000 yr0.1Myr1 Myr3 Myr010203040506070r (AU)−3−2−101Optical Depth (log10)Optical depth contributed by dust (3Myr)J = 1-02-13-24-35-46-5 |
0911.1369 | 1 | 0911 | 2009-11-06T23:08:28 | Three-Body Capture of Irregular Satellites: Application to Jupiter | [
"astro-ph.EP"
] | We investigate a new theory of the origin of the irregular satellites of the giant planets: capture of one member of a ~100-km binary asteroid after tidal disruption. The energy loss from disruption is sufficient for capture, but it cannot deliver the bodies directly to the observed orbits of the irregular satellites. Instead, the long-lived capture orbits subsequently evolve inward due to interactions with a tenuous circumplanetary gas disk.
We focus on the capture by Jupiter, which, due to its large mass, provides the most stringent test of our model. We investigate the possible fates of disrupted bodies, the differences between prograde and retrograde captures, and the effects of Callisto on captured objects. We make an impulse approximation and discuss how it allows us to generalize capture results from equal-mass binaries to binaries with arbitrary mass ratios.
We find that at Jupiter, binaries offer an increase of a factor of ~10 in the capture rate of 100-km objects as compared to single bodies, for objects separated by tens of radii that approach the planet on relatively low-energy trajectories. These bodies are at risk of collision with Callisto, but may be preserved by gas drag if their pericenters are raised quickly enough. We conclude that our mechanism is as capable of producing large irregular satellites as previous suggestions, and it avoids several problems faced by alternative models. | astro-ph.EP | astro-ph | – 1 –
Three-Body Capture of Irregular Satellites:
Application to Jupiter
Catherine Philpott1, Douglas P. Hamilton1, and Craig B. Agnor2
1 Department of Astronomy, University of Maryland, College Park, MD 20742-2421
E-mail: [email protected]
[email protected]
2 Astronomy Unit, School of Mathematical Sciences, Queen Mary University of London,
London, UK E14NS
E-mail: [email protected]
Submitted to Icarus: October 19, 2009
9
0
0
2
v
o
N
6
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
9
6
3
1
.
1
1
9
0
:
v
i
X
r
a
– 2 –
ABSTRACT
We investigate a new theory of the origin of the irregular satellites of the
giant planets: capture of one member of a ∼100-km binary asteroid after tidal
disruption. The energy loss from disruption is sufficient for capture, but it can-
not deliver the bodies directly to the observed orbits of the irregular satellites.
Instead, the long-lived capture orbits subsequently evolve inward due to interac-
tions with a tenuous circumplanetary gas disk.
We focus on the capture by Jupiter, which, due to its large mass, provides the
most stringent test of our model. We investigate the possible fates of disrupted
bodies, the differences between prograde and retrograde captures, and the effects
of Callisto on captured objects. We make an impulse approximation and discuss
how it allows us to generalize capture results from equal-mass binaries to binaries
with arbitrary mass ratios.
We find that at Jupiter, binaries offer an increase of a factor of ∼10 in the
capture rate of 100-km objects as compared to single bodies, for objects separated
by tens of radii that approach the planet on relatively low-energy trajectories.
These bodies are at risk of collision with Callisto, but may be preserved by gas
drag if their pericenters are raised quickly enough. We conclude that our mecha-
nism is as capable of producing large irregular satellites as previous suggestions,
and it avoids several problems faced by alternative models.
1.
Introduction
1.1. Previously suggested capture models
With discoveries accelerating in the last decade, we now know of over 150 satellites
orbiting the giant planets. About one-third of these are classified as regular, with nearly
circular and planar orbits.
It is thought that these satellites are formed by accretion in
circumplanetary disks. The majority of the satellites, however, are irregular and follow
distant, highly eccentric and inclined paths. It is widely believed that irregular satellites
originated in heliocentric orbits and were later captured by their planets, but the details of
how this occurred are still uncertain. At least seven different models have been proposed,
involving dissipative forces, collisions, resonances, and three-body effects. Each model has
its own strengths and weaknesses.
In one long-standing theory, planetesimals are slowed as they punch through the gas
disk surrounding a young, growing planet (Pollack et al., 1979). For this mechanism to be
– 3 –
efficient, the gas must be sufficiently dense to capture the planetesimals in one pass. This
is problematic, however, because if the gas disk does not rarefy substantially in ∼100-1000
years, the orbits of the new satellites will decay inward, leading to collisions with the planet
and its regular satellites. Furthermore, the atmospheres of Uranus and Neptune have only
a few Earth-masses of hydrogen and helium at present, so their gas disks could not have
been as extensive or long-lived as those of Jupiter and Saturn. A likely outcome of this
model, then, is that satellite capture should have been different at Jupiter and Saturn than
at Uranus and Neptune; however, current observational estimates suggest equal efficiencies
(Jewitt & Sheppard, 2005). With a model similar to that of Pollack et al., ´Cuk & Burns
(2004a) found that Jupiter's largest irregular satellite, Himalia, would evolve inward to its
current orbit in 104 − 106 years. This tenuous gas, however, may make capture difficult.
In another model, planetesimals are captured when the mass of the planet increases
(Heppenheimer & Porco, 1977). This mass growth causes the planet's escape velocity to
increase, rendering a previously free planetesimal bound to the planet. For this method
to be effective, the planet's mass must increase substantially on ∼100-1000-year timescales.
However, in most planet formation models (e.g. Pollack et al., 1996), giant planet growth is
hypothesized to take place on timescales many orders of magnitude longer than required by
this capture scenario. Furthermore, Uranus and Neptune's gas deficiency implies that their
growth was of very short duration. Thus, our current understanding of planetary formation
makes this model improbable.
The observation that the four giant planets contain approximately the same number
of irregular satellites (accounting for observational biases; Jewitt & Sheppard, 2005) has
led to a renewal of interest in capture theories that do not depend strongly on the planet's
formation process.
In one such scenario, a planetesimal collides with a current satellite
or another planetesimal in the vicinity of the planet, resulting in its capture (Colombo &
Franklin, 1971). Though collisions were certainly more common in the early Solar System
than they are today, if they resulted in enough energy loss to permit capture, they would
likely also have catastrophically disrupted the bodies. Nevertheless, the fragments might
then have become independent satellites.
A fourth suggestion involves the possible instability in the orbits of the outer planets
early in the Solar System (e.g., by a 2:1 resonance crossing between Jupiter and Saturn).
Outlining the theory of the Nice model of Solar System evolution, Tsiganis et al. (2005) have
shown that such an event could cause Uranus and Neptune to have many close approaches
with each other and with Jupiter and Saturn. During these encounters, the influence of the
massive interloping planet can cause planetesimals to be stabilized as satellites (Nesvorn´y
et al., 2007). This method is promising but has an important disadvantage in that Jupiter
– 4 –
(and Saturn, to a lesser extent) sustains very few close encounters relative to the ice giants.
Thus the gas giants are inefficient at capturing satellites in this way (Nesvorn´y et al., 2007).
Astakhov et al. (2003) examined low-energy orbits that linger near Jupiter and Saturn.
While these bodies are not permanently captured, the authors found that some of them were
stable for thousands of years, long enough to allow a weak dissipative force such as gas drag
to complete the capture process. However, the overall percentage of temporary captures
that do not escape is small, and many of these bodies are threatened by collision with the
planets' large outer satellites (e.g., Callisto and Titan).
Agnor & Hamilton (2006a) examined the capture of Triton from an exchange reaction
between a binary pair and Neptune. Their motivation stemmed from the newly-discovered
abundance of binaries in small-body populations. Currently, it is estimated that binaries
account for ∼30% of Kuiper belt objects (KBOs) with inclinations < 5◦, ∼5% of the rest
of the KBOs (Noll et al., 2008), and ∼2% of large main belt asteroids (diameters > 20
km; percentage increases for smaller objects; Merline et al., 2007). In Agnor & Hamilton's
capture model, a binary is tidally disrupted and one of its members, Triton, is captured as
a satellite. This process is most effective for large satellites like Triton, with radius 1350
km. However, the largest of the other irregular satellites are more than 10 times smaller
than Triton: Himalia at Jupiter is ∼85 km in radius, Saturn's largest irregular, Phoebe, is
∼110 km, Uranus's Sycorax is ∼80 km, and Neptune's Halimede and Neso are only ∼30 km
each. Capturing these satellites via binary exchange reactions would be significantly more
difficult, as we will discuss further below.
Finally, Vokrouhlick´y et al. (2008) examined binary exchange reactions during the first
100 Myr after an assumed Jupiter/Saturn 2:1 resonance crossing, using results of the Nice
model (Tsiganis et al., 2005) to guide their initial conditions. Because planetesimal speeds
relative to the planets are high after the scattering phase of the Nice model, they found that
captures from binaries during that time do not match current orbital parameters and occur
too infrequently to account for today's populations.
1.2. Our model: Capture from 100-km binaries
All of the above models have promising aspects coupled with important limitations. In
this work, we seek to combine the best features of several models into a viable capture sce-
nario. In particular, we examine binaries (as in Agnor & Hamilton, 2006a and Vokrouhlick´y
et al., 2008) as a way to augment capture from low-velocity orbits resulting from three-body
interactions like those studied by Astakhov et al. (2003). While Vokrouhlick´y et al. (2008)
– 5 –
studied exchange reactions in the context of an assumed initial planetesimal population,
we focus on assessing the viability of the mechanism itself. Our goal is to determine how
various parameters of the model affect its plausibility. We examine its viability at Jupiter,
as a number of the above models suggest that capturing at the largest gas giant is especially
difficult.
As the largest of the existing irregular satellites are ∼80-110 km, capture of objects
in this size-range is particularly interesting. Since it is likely that the irregular satellite
population contains collisional families (Nesvorn´y et al., 2003; Sheppard & Jewitt, 2003),
it may be the case that only the largest objects were captured, while the smaller satellites
formed later, via collisions. For this reason, we focus our investigation on capturing the
∼100-km progenitors.
In order to stabilize and shrink the resulting capture orbits, a dissipation source is
required; we suggest a tenuous version of the gas drag originally proposed by Pollack et
al. (1979). Two of Jupiter's irregular satellites, Pasiphae and Sinope, as well as Saturn's
satellite, Siarnaq, and Uranus' Stephano, are found in resonances that seem to require just
such a weak dissipative force (Whipple & Shelus, 1993; Saha & Tremaine, 1993; ´Cuk &
Burns, 2004b).
Furthermore, a tenuous circumplanetary disk is consistent with current theories of late-
stage planetary formation. Jupiter's massive gaseous envelope of hydrogen and helium ne-
cessitates that it formed in the Solar System's circumstellar gas disk. Before the end of its
accretion, Jupiter was likely able to open a gap in the local density distribution of the gas
(for a review, see e.g. Papaloizou et al., 2008). After gap opening, gas continues to leak
into the planet's Hill sphere through the L1 and L2 points, but at a rate much reduced in
comparison to the previous epochs. A tenuous circumplanetary gas disk results (e.g., Lubow
et al., 1999; D'Angelo et al., 2003; Bate et al., 2003), from which material may condense
and regular satellites may accrete near the planet (e.g., Canup & Ward, 2002; Mosquiera &
Estrada, 2003).
In ´Cuk & Burns' study (2004a) of the Himalia progenitor's orbital evolution, they
considered circumjovian nebular conditions consistent with hydrodynamical simulations of
Jupiter's gap opening in a circumstellar gas disk (e.g., Lubow et al., 1999) and found that
the post-capture timescale for evolving this progenitor to its present orbit to be roughly in
the range of 104−106 years. This is similar to the timescale in which extrasolar circumstellar
disks transition from optically thick to thin (∼ 105 years; Skrutskie et al., 1990; Silverstone
et al., 2006; Cieza et al., 2007). The similarity of timescales suggests that satellites captured
at the onset of disk dispersal have a good chance of experiencing stabilizing orbital evolution
while also avoiding collision with the planet.
– 6 –
The timescale for binary capture is very short compared to evolution timescales from
a tenuous gas disk. Therefore, we focus our study first on characterizing the effectiveness
of binary capture in the absence of gas. In the following sections, we critically evaluate our
model for capturing irregular satellites from low-mass (∼100-km) binaries. We begin with a
closer examination of the three-body capture process and then explore parameter space with
a large suite of numerical simulations. We then discuss the ability of gas drag to stabilize
post-capture orbits in Section 4.7.
2. Three-body capture process
Binary capture first requires a close approach between a binary pair and a planet. As
the pair approaches the planet on a hyperbolic trajectory, its two components also orbit
their mutual center of mass (CM). Hence, each member's speed with respect to the planet
is a vector sum of its CM speed (vCM ) and its orbital speed around the CM. If the binary
passes close enough to the planet, it will be tidally disrupted. Following Agnor & Hamilton
(2006a), we make an 'impulse approximation' and assume that disruption is instantaneous,
so that the distance at which tidal disruption occurs (rtd) can be estimated as:
rtd ≈ aB(cid:18) 3MP
m1 + m2(cid:19)1/3
,
(1)
where aB is the semi-major axis of the binary, MP is the mass of the planet, and m1 and m2
are the masses of the binary pair. This tidal disruption radius is the distance to the planet
at which the binary's Hill sphere is no longer larger than the binary itself.
As a result of the impulse approximation, we also assume the orbits of the now-separated
components are dictated by their speeds upon disruption. The speed change of one compo-
nent (∆v1) is approximately equal to its orbital speed around the CM:
∆v1 ≈ ±
m2
m1 + m2 (cid:18)G(m1 + m2)
aB
(cid:19)1/2
,
(2)
where G is the gravitational constant. If the speed of either component is below the escape
speed (vesc) when the binary is split, that component will be captured. This is most efficient
if the incoming vCM is only slightly faster than the value needed for escape. (See Fig. 1.)
The separation of the binary (rB = 2aB, for equal-mass pairs on circular orbits) plays a
key role in determining whether a given encounter will result in a capture. From Eq. 2, we
– 7 –
can see that a smaller separation imparts a higher speed change upon disruption, increasing
the probability of capture. However, the separation must be large enough that the binary can
actually be disrupted. Equation 1 indicates that, not surprisingly, a large separation makes
the binary easier to split. The separation that optimizes capture, then, is one just wide
enough that the binary is disrupted. In addition, the tidal radius is important: the speed
change needed for capture (vCM − vesc, the difference between the two horizontal lines in
Fig 1) decreases for smaller rtd. Thus deeper encounters are more likely to lead to captures.
In much of the current work, we consider the simplified case where Jupiter orbits the
Sun along a circle. In this case, the Jacobi constant (CJ ) for the planet-Sun-interloper three-
body problem is a very useful predictor of the interloper's potential for capture, taking on the
role of v∞ from the two-body approximation. Although our model contains four bodies, and
the Jacobi constant is a three-body construct, it is an excellent approximation to consider
the CM of the binary as one body moving in the Sun-Jupiter system up until the point of
disruption. The gravitational energy between the binary components is negligible after they
separate. Thus after disruption, we essentially have two separate three-body problems, one
for each binary component, and we can make use of the Jacobi constant throughout the
entire simulation.
If CJ ≥ CJ,crit, the critical value for capture, bodies in the vicinity of the planet are
bound by so-called zero-velocity curves (ZVCs) that enclose Jupiter and constrain particle
motions (Fig. 2). For Jacobi constants lower than CJ,crit (i.e., higher energies), one large zero-
velocity curve surrounds both Jupiter and the Sun and bodies can enter and exit Jupiter's
Hill sphere freely. The critical Jacobi constant represents the boundary between these pos-
sibilities. Murray & Dermott (1999) give its value: CJ,crit ≈ 3 + 34/3µ2/3 − 10µ/3, where µ
= MP
and M⊙ and MP are the masses of the Sun and the planet, respectively. Here we
use dimensionless units in which G, the Jupiter-Sun distance, and the sum of the solar and
jovian masses are equal to 1. For the Jupiter-Sun system, which is the focus of the current
paper, µ = 9.53 ×10−4 and CJ,crit ≈ 3.0387.
M⊙+MP
Figure 3 illustrates a typical capture involving Jupiter. In the bottom panel, the Jacobi
constant of the binary pair prior to its split is lower than the critical value, meaning that
initially, the binary has too much energy to be bound. The oscillations in the bodies' pre-
disruption CJ are due to gravitational interactions between the binary components. At
the time of disruption (t ≈ 8 yr), one component sharply gains energy (CJ decreases),
while the other component experiences a corresponding energy loss (CJ increases). In this
example, one component's final CJ is higher than the critical value, signifying that it is
permanently bound to Jupiter. Though the Jacobi constant is very valuable when considering
a circularly-orbiting Jupiter, a disadvantage is that it cannot be extended to cases with non-
– 8 –
zero eccentricity. In this paper, we make the simplifying assumption that eJ = 0 (rather
than the true value of ∼0.048) in order to better elucidate important physics of the problem.
In Fig. 4, we plot the orbits of the binary components shown in Fig. 3. Low-velocity
orbits like these are characterized by multiple close passes by the planet (cf. Hamilton &
Burns, 1991). The separation is disturbed by the strong tidal force during each of these
passes, but the binary splits only after it comes within the tidal disruption radius (see top
and middle panels of Fig. 3).
The binary capture mechanism is most effective at producing permanent or long-lived
captures if i) the mutual orbital speed of the binary is high, and/or ii) the encounter speed is
low. Agnor & Hamilton's (2006a) work examined Neptune's moon Triton, which is somewhat
of a special case because it fulfills both of these criteria – its size means that its orbital speed
around a close companion would be high, and typical encounter speeds at Neptune in the
early stages of planet formation are relatively low.
The direct three-body capture mechanism is much less effective for most other irregular
satellites which are ∼100 km or smaller in radius. Furthermore, because of Jupiter and
Saturn's sizes and proximity to the Sun, encounter speeds at the semi-major axes of the gas
giants' irregular satellites are relatively fast, vCM ≈ 3 km/s. To produce a large enough
energy change for capturing directly to the current satellites' locations, binary components
must be orbiting each other at speeds comparable to their encounter speeds. This would
require binary companions of order Mars- or Earth-sized (Agnor & Hamilton, 2006b) – an
uncommon occurrence even in the early Solar System.
Accordingly, in this work, we relax the requirement that moons are captured directly
to their present orbits. In the example discussed above (see Fig. 4), the final orbit of the
captured satellite extends almost to the Hill radius (rH ), whereas the actual satellites at
Jupiter are significantly more tightly bound. We investigate the idea that the objects were
first captured to these distant orbits, and a subsequent period of orbital evolution (e.g., by
weak gas drag) led them to their current configurations.
The post-capture evolution is a key component in our model because it allows for capture
from small binary pairs, even though they deliver satellites to very distant orbits. Binaries
with primaries of order 100 km were certainly much more numerous than those with planet-
sized primaries, even in the early Solar System. Models that rely on gas drag for capture (e.g.
Pollack et al., 1979) require both i) dense gas (to enable capture) and ii) rapid dispersal (to
prevent satellite loss to the planet). By contrast, our model requires no gas for capture and
puts only weak constraints on gas required for orbital evolution. In particular, we require
only that the product of the gas density and its residence time around Jupiter be large
enough that the requisite amount of evolution can occur.
– 9 –
3. Numerical model
The goal of this work is to characterize the overall effect of binaries on the probability
of capturing bodies on planet-crossing paths. We focus primarily on captures at Jupiter,
which has the most irregular satellites and has many sources of small bodies nearby. Also,
as discussed above, capture at Jupiter has shown to be difficult, especially because of its
large size and fast encounter speeds for approaching bodies. Thus these simulations provide
the most stringent test of our model.
Our integrations include the Sun, Jupiter, and a binary or single object, in a planet-
centered frame. In order to examine binaries' effectiveness at producing long-lived captures,
we compare them to captures of single-body interlopers. While only tidally disrupted binaries
can be captured permanently, unbound single bodies can remain near the planet for long
periods of time (e.g. Astakhov et al., 2003). We define a 'capture' to be a body that remains
near Jupiter for 1,000 years, the duration of each simulation. Furthermore, we define 'binary
capture' to mean capture of one or both of the bodies that originated together as binary
components. Note that these definitions encompass both permanent (energetically bound)
and long-lived temporary captures. With captures of single bodies as a baseline, we are able
to measure the enhancement due to binaries.
Our simulations are performed with HNBody, a hierarchical N-body integration package,
and HNDrag, a companion code for applying non-gravitational forces to the particles and for
detecting close approaches (Rauch & Hamilton, 2002). For most of this work, we use only
the close-approach detecting capabilities of HNDrag and include only gravitational forces.
We use HNBody's Bulirsch-Stoer integrator with a specified accuracy of one part in 1014.
The adaptive-stepsize Bulirsch-Stoer integrator is much more efficient than a symplectic
integrator here because while a small stepsize is needed initially to resolve the orbital motion
around the binary CM, it can be greatly increased after disruption.
We ran about 200 sets of three- or four-body simulations examining a range of Jacobi
constants for each interloper (2.95-3.037), as well as varying the binaries' radii (65-, 100-,
and 125-km), and separations (1-1000 body radii). For each set of parameters, we generated
10,000 binaries or single objects, for a total of ∼2×106 simulations, each following the bodies
for 1,000 years. We started all of the interlopers of a given set at the same distance from
the planet, ranging from 1.0 - 1.4 rH. (Section 4.4 contains a discussion of the effects of
starting distances on capture statistics.) The choice of Jacobi constant and starting distance
– 10 –
constrains the possible initial positions of the binary CM. Fig. 2 shows Jupiter's Hill sphere
overplotted with ZVCs corresponding to the Jacobi constants that we studied. Bodies are
energetically unable to cross their zero-velocity surfaces, and thus starting with, say, CJ =
3.037 at 1.0 rH from the planet restricts the body's initial position to the two small 'endcaps'
of the Hill sphere along the Jupiter-Sun line. For smaller CJ , these allowed areas are larger
and finally encompass the entire Hill sphere for CJ .3.025. The bodies' initial speeds are
also constrained by the specified Jacobi constant, and we choose the velocity to point in
random inward directions.
For simplicity, we set the binary components to orbit each other on circles and the
binary angular momentum to be perpendicular to Jupiter's equatorial plane. We ran each
initial condition with five different binary orbital phases (equally-spaced mean anomalies)
and averaged all capture statistics over the five phases. Throughout the simulations, we
monitored the bodies, weeding out very close approaches between any two objects and noting
each body's close approaches to Jupiter. (Collisions between binary members do occur, but
these are rare and of limited interest, since the merged object simply behaves as a single
interloper with the same CM speed.) To shorten the computational time required, we stopped
integrations in which all of the incoming objects traveled further than 2-5 Hill radii from
Jupiter, depending on the bodies' starting distance from the planet.
4. Results
4.1. Relationship between inclination and CJ
In our simulations, we find that the inclination of the approach trajectory is correlated
with the initial Jacobi constant, which is helpful in providing physical intuition for the
meaning of CJ. This correlation was first noticed numerically by Astakhov et al. (2003); here
we confirm their finding numerically and provide an analytical explanation. Fig. 5 displays
this CJ -inclination relationship. We see a clear correlation of CJ with mean inclination:
lower Jacobi constants are indicative of retrograde orbits, while prograde orbits have larger
CJ . For clarity, the plot shows only a representative population of bodies: 100-km binary
members that result in a capture, with inclinations calculated at the closest approach of
each body's first pass by Jupiter. However, the relationship holds for all close approaches of
binaries or single objects, captured or not. This plot provokes two main questions: what is
the physical cause of this trend and why is there such high scatter in inclination at a given
Jacobi constant? We address the question of scatter first.
One complexity in making this plot is that all orbital elements including inclination are
– 11 –
poorly defined at large distances from Jupiter, as solar tides are comparable to Jupiter's
gravity at the Hill sphere. Accordingly, we were careful to calculate the inclinations only
at orbital pericenter where solar tides are weakest so that inclination is always well defined.
Poorly-defined orbital elements, therefore, are not the source of the scatter. Furthermore,
the variations look nearly the same when we plot single objects rather than binaries, which is
expected since disrupting the binary results in an energy change that only slightly alters CJ
(e.g., Fig. 3). Finally, the scatter is present even when we consider one individual object's
multiple pericenter passages rather than those of an ensemble of objects. Thus the spread
in inclination is real and is due to the response of a single captured object to the solar tidal
force.
The scatter in inclination as well as the inclination-CJ trend can be understood analyt-
ically by writing the Jacobi contant in terms of planetocentric orbital elements rather than
the heliocentric orbital elements used in deriving the standard Tisserand constant (Murray
& Dermott, 1999). We begin with the planet-centered 'generalized Tisserand constant' de-
rived in Hamilton & Krivov, 1997 (their Eq. 4) and neglect the solar tidal term since it
is complicated and unimportant at pericenter where we measure inclination. We then non-
dimensionalize the equation as described in Section 2, and finally, to conform to standard
usage (Murray & Dermott, 1999), we add the constant 3 to the final result and find:
CJ
′ = 3 +
31/3µ2/3
¯a
"1 + 2(cid:18) ¯a3(1 − e2)
3
(cid:19)1/2
cos(i)# ,
(3)
where µ is the mass ratio as defined above; ¯a = a/rH; and a, e, and i are the captured satel-
lite's semi-major axis, eccentricity, and inclination, respectively. Because we have neglected
the solar tidal term and used planetocentric orbital elements, this expression is valid only
near the planet where solar perturbations are weak.
For close orbits of the planet (¯a << rH), orbit-averaging the effects of the tidal force
shows that the semimajor axis ¯a remains constant. Accordingly, Eq. 3 leads directly to
the Kozai constant, K = √1 − e2 cos(i). This constant explains the coupled oscillations
in eccentricity and inclination that characterize the Kozai resonance. If K were precisely
conserved, orbits would not be able to switch between i < 90◦ (which have K > 0) and
i > 90◦ (which have K < 0). We do, however, see such prograde-to-retrograde transfers
in Fig. 5, which indicates that, as expected, K (and therefore ¯a) is not constant for our
distant orbits (see, e.g., Hamilton & Burns, 1991).
In addition, at apocenter where the
solar tidal force is strongest, the orbital elements themselves are poorly defined and Eq. 3
is only approximate. Thus between each pericenter passage, the orbital elements (including
inclination) are scrambled by the solar tidal force leading to dispersion like that seen in
– 12 –
Fig. 5.
The trend observed in Fig. 5, decreasing inclination for increasing Jacobi constant, is
neatly explained by Eq. 3. Testing the equation quantitatively, we estimate ¯a = 0.5 and e =
0.7 for a typical orbit of a body captured at Jupiter (e.g., Fig. 4). A purely prograde orbit
′ = 3.020. These
(i = 0) gives CJ
values roughly correspond to the range of Jacobi constants seen in Fig. 5, despite the rather
large approximations that we have made. The inclination-CJ correlation is strong enough
that we will often use the term prograde to refer to orbits with CJ ∼ 3.03, and retrograde
to mean CJ ∼ 3.01.
′ = 3.036, while a purely retrograde orbit (i = 180◦) gives CJ
4.2. Modes of capture
For each binary-planet encounter, there are four possible outcomes: (1) neither com-
ponent captures (hereafter known as '0C'), (2) one component captures ('1C'), (3) both
components capture together as an intact binary without splitting apart ('2C-BIN'), or (4)
the binary is disrupted and both components capture individually ('2C-IND'). The frequency
of each type of outcome depends on the characteristics of the binary. Fig. 6 shows the out-
comes that result in a capture (i.e., 1C, 2C-BIN, and 2C-IND) for 65-km binary pairs with
initial CJ = 3.037, as a function of the initial separation, rB, of the binary. The separation
can be altered significantly prior to disruption during close approaches to the planet. The
number of captures for a set of single objects is also plotted for comparison; these must be
temporary captures since there is no energy loss.
The rate of 2C-BIN captures is largest when the separation is small (and thus the com-
ponents are tightly bound to each other). For small-enough separations, the tidal disruption
radius is so close to the planet that very few binary orbits cross it (see Eq. 1). Here, most
of the binaries remain intact, and the 2C-BIN rate nearly matches that of single objects.
When we increase the binary separation, more binaries are split, and the 2C-BIN capture
percentage monotonically drops to zero, as expected.
Disrupting binaries leads to more possibilities for capture of individual objects. Accord-
ingly, as the separation increases, the 1C and 2C-IND capture rates rises from zero. For
the 1C population of 65-km objects with CJ = 3.037, there is a peak in capture efficiency
of ∼5 times that of a single body at a separation of ∼20 body radii (RB). This separation
represents the optimal balance between disrupting a high percentage of the binaries and de-
livering the most energy upon disruption. The optimum separation varies depending on the
mass of the binary. At larger separations, the binding energy decreases, leading to smaller
– 13 –
energy kicks and a diminished capture rate.
The 2C-IND percentage has a peak at the same separation as the 1C group. These
binaries likely split during orbital phases where the energy is distributed almost equally
between the components. The number of 2C-IND captures is never more than a few percent
of the 1C captures, but the two populations peak at rB ∼ 20 RB for the same reasons. More
widely separated binaries are disrupted with a smaller energy change. Because of this, the
two components are more likely to have similar energies and post-disruption fates, causing
an increase in 2C-IND captures at larger separations.
Unlike the case for 1C and 2C-IND capture where energy is lost and capture can be
long-lived or even permanent (as in Fig. 3), capture of singles or intact binaries (2C-BIN)
is necessarily temporary. This could be an advantage for 1C captures, which have more
stable, lower-energy initial capture orbits. The details of the final comparative satellite
yields depend on the subsequent orbital evolution, which is determined by the gas present
at the time of capture and its dissipation timescale.
4.3. Effects of binary mass and orbital separation
Having explored the physical meaning of CJ and the possible types of captures, we now
discuss the results of the numerical simulations. In this section, we consider cases of equal-
mass binaries encountering Jupiter with the planet on a circular orbit, and we examine the
effects of the bodies' masses, binary separations, and initial Jacobi constants. We performed
integrations over a range of Jacobi constants: 2.95 ≤ CJ ≤ 3.037, where CJ ≈ 3.0387 is the
critical value above which transfer orbits between Jupiter and the Sun are impossible (see
Fig. 2). For CJ ≤ 2.99, no captures resulted for any of the parameters we tested, although
capture at these low Jacobi constants could certainly occur for larger-mass binaries. For
now, we consider the fate of bodies started from the Hill sphere (following Astakhov et al.,
2003); in Section 4.4, we discuss the importance of alternative starting distances.
We examined masses corresponding to pairs of objects each with radii 65-km, 100-km,
and 125-km (assuming a density of ∼2 g/cm3). Fig. 7 displays the results of these mass
studies. We see that capture rates increase for higher masses: the 125-km capture rate is
slightly higher than the 100-km rate throughout the range of Jacobi constants, and they
differ most significantly from 65-km binary pairs for CJ > 3.03.
For single objects, mass has no effect on capture probability, but for binaries, larger
total mass leads to more rapid orbital speeds and a higher speed change upon breakup. This
can be seen by eliminating aB from Eqs. 1 and 2, with mass ratio ( m2
) and tidal distance
m1+m2
– 14 –
(rtd) held constant; the result is ∆v1 ∼ (m1 + m2)1/3. Accordingly, larger masses generally
lead to increased capture rates.
Another important result is that capture rates from binaries are extremely sensitive to
the binary's separation, rB = 2aB. For each of the masses we examined, we determined the
optimum separation of the binary required to achieve the maximum capture probability. We
used Eqs. 1 and 2 to guide our separation selection, and we integrated each point on Fig. 7
with several different binary separations to determine the optimal value.
In Fig. 7, we have plotted statistics using a single separation for each mass over the
range of Jacobi constants. For most of the Jacobi constants studied for a given mass, the
optimal separations are very similar, ∼10 RB. An important exception is for the highest CJ
value tested, 3.037 (recall that this corresponds to mostly prograde encounters – see Fig. 5),
which had maximum captures at a larger separation (∼20 RB) than the typical optimal
value. Because we have not plotted this optimal value in Fig. 7, the curve declines sharply
at CJ = 3.037. As is clear from Section 4.2, optimizing the separation makes a significant
difference in capture rate, especially for binaries whose CJ values are close to the capture
threshold. Changing from ∼10 RB to the optimal 20 RB for CJ = 3.037 increases the capture
percentage from ∼2% to ∼10% (100-125 km objects) and ∼1% to ∼3% (65-km, see Fig. 6).
Binary capture rates also appear to depend strongly on initial Jacobi constant. It is
tempting to compare capture efficiencies for low and high Jacobi constants (retrograde and
prograde orbits). A direct comparison of these rates, however, cannot be made for reasons
that will become apparent in the next section. We can, however, compare the binary statistics
to those of single objects. The bodies that originate in binaries capture with similar rates as
the single objects below CJ ≈ 3.015, but as CJ increases, the effects of the binaries become
more visible, rising by an order of magnitude in efficiency at delivering objects to Jupiter.
Part of the reason for this is probably that the retrograde binaries are harder to split than
progrades because of their orientation to the planet. Another explanation, particularly for
the highest initial Jacobi constants, is that these encounters are close to the critial energy
barrier for capture, and so the energy change from disruption of the binary is more likely to
result in capture.
4.4. Effects of starting distance
4.4.1. Contamination from bound retrograde orbits
Thus far we have discussed results from initial conditions that launch objects from the
Hill sphere. In the course of this study, we discovered that the choice of starting distance can
– 15 –
significantly affect the capture statistics. While the Hill sphere is defined as a rough stability
boundary beyond which the Sun's gravitational influence is stronger than the planet's, in
practice, stable retrograde orbits can extend out to distances slightly beyond the Hill radius
(see, e.g., Hamilton & Burns, 1991). In contrast, stable prograde orbits (e.g., Fig. 4) are
always well within the Hill sphere. Starting the integrations with bodies on the Hill sphere,
then, risks starting on an already-stable retrograde orbit. This causes ambiguity in the
capture statistics – which orbits were always near Jupiter and which truly came in from
infinity?
We demand that true captures originate in heliocentric orbit and transition to planet-
centered orbits, remaining for at least 1,000 years. To differentiate between these and mis-
leading 'captures' from bodies that are already orbiting the planet at the beginning of our
integration, we took the single-body capture orbits and integrated the initial conditions back-
wards in time for 1,000 years. We then separated the captures that came in from infinity
and those that were always present near the planet. We found that when starting on the
Hill sphere, most of the resulting retrograde captures of single objects at low CJ had always
been orbiting the planet and were in fact not true captures.
We examined several other launch distances and determined that starting from 1.1 Hill
radii (rH) and beyond eliminates nearly all contamination from already-stable retrograde
orbits. Figure 8 compares the capture rates for single objects beginning from 1.0 rH and
1.1 rH. We see that while the high-CJ captures are uncontaminated, launching from 1.0 rH
for CJ < 3.03 leads to many false captures. In fact, over one-third of the 1.0 rH captures
are contaminations. (This fraction is much smaller for binaries, where there are many more
prograde captures than retrograde.) When the objects are launched from 1.1 rH , we see no
artificial captures at all. Similar tests at 1.2 rH and 1.4 rH also show no false captures. Al-
though stable orbits do exist to these and larger distances (Henon, 1969), they are apparently
extremely rare.
4.4.2. Scaling to different starting distances
So, is it safe to compare prograde and retrograde capture statistics for starting distances
beyond 1.1 rH? Starting outside the Hill sphere results in a decrease in true captures over the
full range of CJ (as seen in Fig. 8). This is expected, because fewer of these bodies experience
close approaches with the planet. However, prograde captures are more drastically reduced
than retrogrades. Determining the reason for this is critical to answering our question.
At high Jacobi constants (& 3.03, which correspond to mostly prograde orbits), the
– 16 –
binary's ZVCs are closed around the planet except for only a small neck on either side of
Jupiter, toward and away from the Sun (see Fig. 2). In contrast, lower Jacobi constants
(and thus retrograde orbits) allow for more freedom of movement in the region between
the Sun and Jupiter. Retrograde orbits can approach the planet from all directions, while
progrades are limited to the entering via the narrow ZVC necks. Starting further than the
Hill sphere means that a smaller fraction of the initial prograde trajectories pass through
the zero-velocity necks and approach Jupiter, resulting in a decrease in captures. We tested
the hypothesis that the shape of the prograde ZVCs is the primary reason for the decrease
in prograde captures with the following procedure:
1) Determine the capture rate at 1.2 rH.
2) Determine the crossing rate of 1.2 rH orbits inside 1.1 rH.
3) Scale the 1.1 rH capture rate by the crossing rate from step (2).
We use the statistics for 1.1 rH and 1.2 rH because they are free from any contamination
from the false retrograde captures discussed above. The result of step (3) is what we expect
the 1.2 rH capture percentages would be if the ZVC shapes are the reason for the decline
in prograde captures from 1.1 rH to 1.2 rH. Comparing these scaled capture rates with the
actual 1.2 rH percentages (as shown in Fig. 9), we see that this scaling accounts for most
of the difference between the two capture rates, to within 20%. This indicates that our
prediction is valid.
In principle, these results also account for the reduction in true retrograde captures, but
the number statistics are so low for retrogrades at these launch distances that the actual and
scaled capture percentages are equivalent to within the error.
We find, then, that because of the differing geometries of their ZVCs, we cannot directly
compare prograde and retrograde statistics. This is true even when starting beyond the limit
for already-stable retrograde orbits, chiefly because prograde orbits are so sensitive to their
initial distance. To circumvent this effect, we would have to start far enough away from
the planet that further increasing the starting distance would result in an equal fractional
decrease in captures for retrogrades and progrades – in other words, where the geometry of
the prograde ZVCs is no longer the dominant reason for the decrease. Our simulations, out
to 1.4 rH , did not reach this threshold and were limited by number statistics. We anticipate
that comparing prograde and retrograde captures would require starting still further from
the planet than our trials, and would demand much larger initial populations.
– 17 –
4.4.3. Starting from 1.1 rH vs. 1.0 rH
Launching binaries from 1.1 rH rather than 1.0 rH results in an overall decrease in
captures for the reasons described above, but many of the characteristics of the capture
orbits remain relatively unchanged. For example, the inclination distribution for the 1.1 rH
captures is almost identical to the 1.0 rH population plotted in Fig. 5. Also, comparing
Fig. 9 with Fig. 7, we see that the trend in the capture percentage with CJ of the 100-km
binaries at 1.1 rH and 1.0 rH are generally similar.
Figure 10 displays the modes of capture vs. Jacobi constant for 100-km binaries and
single objects starting from both 1.0 rH and 1.1 rH. Starting at 1.1 rH results in a depletion
of 2C captures relative to 1C captures. This is because many of the retrograde 2C-IND and
2C-BIN captures were likely the artificial retrogrades which are eliminated when starting
from 1.1 rH . The single captures also appear depleted relative to the other curves, but in
fact, the binaries are diminished by a roughly comparable amount. Overall, there are few
qualitative differences between the 1.0 rH and the 1.1 rH cases.
We note that other groups (including Astakhov et al., 2003 and Astakhov & Farrelly,
2004) have initiated bodies from the Hill sphere without considering how this affects the
resulting statistics, most notably that many of their retrograde encounters are started on
already-stable orbits. For future studies of capture near Jupiter, we recommend the following
procedure: 1) generate bodies starting from 1.0 rH , 2) integrate each body backwards in
time, 3) eliminate those that remain near the planet when integrated backwards, and 4)
integrate the remaining bodies in normal time direction. The benefits of this approach over
simply starting from 1.1 rH are that it eliminates all stable-retrograde contamination while
preserving high number statistics. Note, however, that this procedure still does not allow
a valid comparison between prograde and retrograde statistics. To obtain the true ratio of
captured progrades and retrogrades, it would probably be best to start the interlopers on
heliocentric orbits.
4.5. Scaling to unequal binary masses
The impulse approximation (Eqs. 1 and 2) implies that a binary component's likelihood
of capture depends only on 1) the binary's tidal disruption radius and 2) the component's
instantaneous speed at the time of disruption. Accordingly, for any mass ratio m1 : m2
and semi-major axis aB, we can find an equal-mass binary with the same rtd and the same
component speeds as m1 and another equal-mass binary that matches these values for m2.
Setting rtd and v equal for the two mass ratios, we solve for the component mass, m′, and
– 18 –
the semi-major axis, aB
′, of the equal-mass binary matching m1:
and
m′ =
3
4m2
(m1 + m2)2
aB
′ = aB(cid:18) 2m2
m1 + m2(cid:19) .
(4)
(5)
The problem is symmetric, so for the equal-mass binary matching m2, we simply exchange
m1 and m2 in the above equations.
We tested these predictions for 3:2 and 4:1 mass ratios and display the results in Table
1. Each of the equal-mass components captures with the same efficiency (to within 10%)
as either m1 or m2. This is strong validation of the impulse approximation. This sort of
scaling requires, of course, that the binary be split and therefore applies only to our 1C and
2C-IND results. By contrast, undisrupted binaries (2C-BIN) behave as single objects and
have a capture efficiency that is independent of mass.
The impulse approximation allows us to make other predictions as well. For example,
we can reverse the above scenario and predict the capture rates of two equal-mass binaries
corresponding to one unequal-mass pair for which the capture rates of each component are
known.
Further, given the capture rate for a single equal-mass binary, we can predict the capture
efficiency of one component of an unequal-mass pair, if one parameter of the unequal-mass
binary (m1, m2, m1 + m2, or aB) is set. This means that each equal-mass pair matches to a
whole family of unequal-mass binaries.
Finally, we can extend these techniques to guide large-scale simulations. Using the im-
pulse approximation, and given equal-mass capture statistics for all combinations of relevant
masses and separations, we can predict the capture rates for all unequal-mass binaries – with
any mass ratio, total mass, and separation. Similarly, knowledge of unequal-mass capture
rates for one component of a binary with a fixed mass ratio allows us to estimate the capture
efficiencies of all equal-mass binaries.
Scaling with the impulse approximation is a powerful way to predict capture rates.
Practically speaking, it means that capture of any unequal-mass binaries can be predicted
by studying just equal-mass cases, and accordingly, we have restricted our numerical studies
to binaries with equal masses.
– 19 –
4.6. Jupiter's eccentricity
Most scenarios, including our own, propose that irregular satellite capture occurred
early in the Solar System's history (Section 1.1). It is likely that Jupiter's eccentricity was
closer to zero at this time and its current value was obtained later. Thus we have chosen to
use eJ = 0 in our simulations and believe it is a reasonable assumption.
How would Jupiter's current eccentricity affect capture? It is not straightforward to
extend this study to a non-zero jovian eccentricity for the primary reason that the Jacobi
constant is no longer a constant of the motion. Though Jupiter's current eccentricity is small
(eJ ∼ 0.048), it causes the calculated CJ of an interloper to vary by ∼0.2 over an orbit.
This is a variation about ten times larger than our entire range of tested Jacobi constants.
Thus we cannot simply assign an initial CJ to binaries or single objects approaching an
eccentric Jupiter and compare the capture results with the equivalent circular case. Only for
eccentricities less than a hundredth of Jupiter's would the errors introduced in the Jacobi
constant be acceptable.
The results of Astakhov & Farrelly (2004) suggest an order of magnitude lower capture
probability for single objects in the eccentric case as compared to capture by a planet orbiting
on a circle. However, their initial conditions were generated using the same Jacobi constant
values in the eccentric and the non-eccentric cases. Because of the variation in CJ induced
by eccentricity, their method actually produces a much larger range of Jacobi constants (and
much higher approach speeds) for tests with an eccentric planet, and the results should not
be directly compared with the circular case. We believe their conclusions give artificially low
eccentric-case capture rates for this reason.
We do not expect Jupiter's eccentricity to strongly affect capture. As the timescale for
capture is much shorter than Jupiter's orbital period, the planet's instantaneous location is
the most relevant factor. At pericenter, Jupiter's Hill sphere is slightly smaller than at its
average distance from the Sun and encounter speeds are slightly higher, likely resulting in
fewer captures. The opposite can be expected at apocenter, causing the effects of eccentricity
to average out and probably produce little change in overall capture statistics. To truly
know the effects of the planet's eccentricity would require large-scale integrations with the
interlopers originating on heliocentric orbits. This is beyond the scope of our current work,
but again, we suspect the difference in capture rates will be small.
– 20 –
4.7. Survivability of captured objects
The post-capture orbits are initially very irregular and prone to collisions with Jupiter's
Galilean satellites (particularly its outermost, Callisto) or the planet itself. Figure 11 displays
the percent of captured bodies that are delivered onto orbits that do not cross Callisto, which
orbits Jupiter at 26 RJ .
An interesting feature is the lack of any captures without close approaches for CJ =
3.02 and 3.025. These Jacobi constants correspond to orbital inclinations near 90◦ (see
Section 4.1), where the Kozai effect is strongest. The Kozai mechanism causes the orbits to
become highly eccentric and subject to collision with Jupiter or one of the Galilean satellites.
This is the primary reason that no existing satellites in the Solar System have inclinations
near 90◦.
All of the surviving binaries plotted here capture as 2C-BIN (though not all 2C-BIN
captures are survivors), largely because of the separations we have studied. The separations
of these binaries are optimized to give maximum capture percentages, leading to tidal dis-
ruption radii that are very close to Jupiter (Eq. 1). Thus these surviving binaries, which
by definition have closest approaches outside Callisto, do not cross rtd and remain intact.
The separations we have used, which are optimal for capture, do not correspond with the
maximum survival rate. Somewhat larger separations would result in a higher percent of
surviving captures, with a lower percent of overall captures.
The consequences of this are apparent in Figure 11. For CJ < 3.02, the curves for single
objects and all three binary masses are equivalent to within the error. This is because most
of the binary captures in this CJ range are 2C-BIN. Since 2C-BIN captures act as a single
entity with no alteration from tidal disruption, the captured population for CJ < 3.02 is
similar to that of the single bodies and the fraction that avoid Callisto is also similar. While
the surviving binaries for CJ > 3.025 are still all 2C-BIN, 1C captures dominate the 2C-BIN
captures in this Jacobi constant range, but none of the 1C captures are safe from Callisto.
Therefore, the percent of the total binary captures that survive at high Jacobi constants is
very small.
Thus we find that while retrograde captures (low CJ ) are rare (see Fig. 7), most of them
are safe from collision with Callisto. On the contrary, prograde captures from binaries (high
CJ ) are numerous but prone to collision. This is bad news for binary capture – the largest
enhancements occur for captures that are most likely to be removed by interactions with
Callisto. As we have already discussed, examining larger separations will lead to a larger
percentage of capture orbits that are safe from collision. This may be a significant effect.
How else can collisions be avoided?
– 21 –
One possible way out is if the captures occurred before Callisto was formed. At these
early times, a dense accretion disk surrounded Jupiter, and the strong gas drag could have
augmented capture rates (as in Pollack et al., 1979). However, this is not compelling, because
satellites captured at this time would be prone to loss by orbital decay and later by collisions
with forming proto-satellites.
The most likely mechanism for preventing captured bodies from colliding is gas drag
from the remaining gas present outside Callisto's orbit at the time of capture. A small
amount of gas is necessary for our mechanism in order to shrink capture orbits to their
current sizes. This process can also increase the pericenters of the captured objects, causing
them to avoid collision with Callisto. A typical collision timescale for a Callisto-crossing
orbit with a = 0.5rH and i = 10◦ is on the order of 106 years, long enough for gas to evolve
the satellites onto safe orbits (see Section 1.2). The timescale is longer for more tilted orbits,
but significantly shorter for retrograde orbits.
A tenuous gas around Jupiter and Saturn near the end of planet formation is consistent
with our current understanding of planet formation (Section 1.2).
If gas was present at
the time of capture, its structure and density are not well constrained, and thus we do not
focus on the orbital evolution process itself in this work. As a simple example, however,
we simulated a drag force that acts against the velocity vector and show that it is able
to both shrink the post-capture orbits and prevent the new satellites from collision.
In
Fig. 12, we plot the initial and final states for a prograde and retrograde orbit. The gas
drag is applied after capture for simplicity; this is valid because the timescale for temporary
capture is ∼months, and the effects of the gas are negligible over such a short time. The
final states are chosen so that the orbits lie approximately where the current progrades and
retrogrades orbit at Jupiter (at ∼ 1
2rH, respectively), and they have pericenters
outside Callisto's orbit. The drag strengths were set so that the evolution for both orbits
occurs over 25,000 years. However, with a more tenuous gas than in this example, the same
evolution could take place on timescales 10-100 times longer. The binary capture mechanism
discussed here constrains only the amount of orbital evolution, not the evolution rate, and
hence avoids the satellite loss problem of capture-by-gas-drag models.
4rH and ∼ 1
5. Discussion
The new model discussed here, capture from low-mass binaries with subsequent or-
bital evolution, has both significant advantages and disadvantages in comparison to other
suggested models.
– 22 –
One important advantage is that capture is viable at Jupiter and Saturn, unlike three
of the models discussed in Section 1. For example, Agnor & Hamilton's (2006) direct three-
body capture model works only for very large bodies in the gas giants' high-approach-speed
environments. Also, the theory of Nesvorn´y et al. (2003) requires close approaches among the
giant planets, of which Jupiter has very few in the Nice model scenario. Saturn encounters
its outer neighbors more frequently than Jupiter, but it still suffers from few close approaches
overall in this model. In Vokrouhlick´y et al. (2007), the capture statistics from planet-binary
encounters are low for all planets, but especially Jupiter and Saturn. This is primarily
because of the high relative velocities assumed in their model. Though these latter two
capture mechanisms in their current forms cannot explain the gas giants' irregular satellites,
they are worth further study, perhaps in the context of altered versions of the Nice model
or other early Solar System models.
Our model also has an important advantage over that of Pollack et al., 1979, in that our
capture scenario allows for a much weaker gas, since the gas here is needed for shrinking the
orbits, not capturing satellites. Also, the gas in our model can persist for much longer than
that in Pollack et al., as a weaker gas does not have the problem of quickly destroying the
captured bodies. Furthermore, various groups (e.g., Canup & Ward, 2002) have proposed
that the Galilean satellites formed from a tenuous gas; if true, the gas was most likely even
thinner when the irregular satellites were captured. Thus it is important that our capture
mechanism does not rely on a dense gas.
To directly compare our mechanism with that of Astakhov et al. (2003), we need to
consider both relative capture rates (and their survivability) and the prevalence of binaries
vs. single bodies.
In our simulations, we find that binary capture can provide a signifi-
cant advantage over capturing from populations of single bodies for binaries with particular
characteristics: high enough masses (& 100 km), optimal separations (∼10-20 RB), and low
incoming energies (corresponding to Jacobi constants & 3.02 and mostly prograde encoun-
ters). However, like Astakhov et al. (2003), we also find that the probability of captures (from
either binaries or single bodies) of avoiding collisions with Callisto is low for readily captured
progenitors. In Section 4.7, we discussed that this problem can be alleviated by altering the
capture orbits with the surrounding gas or by capturing binaries with larger-than-optimal
separations that do not lead to Callisto-crossing orbits.
So, how common were easily-captured binaries early in Solar System history? This
question is difficult to assess. Observational surveys of the current population of the cold,
classical Kuiper belt (i.e. objects with modest inclinations and eccentricities) find a 30%
binary fraction among bodies larger than 100 km (Noll et al., 2008), many with nearly
equal mass components. Also, recent studies of planetesimal formation have suggested that
– 23 –
large, & 100-km bodies may form quickly in the gaseous proto-planetary disk, providing
the building blocks of subsequent planet formation (Johansen et al., 2007; Cuzzi et al.,
2008). Binary formation is likely to be contemporaneous with the formation of these bodies
(Nesvorn´y, 2008). Further, Morbidelli et al. (2009) have shown that the size-frequency
distribution of asteroids in the main belt is consistent with large initial planetesimals, at
least ∼100-km in size. Together these results indicate that the ∼100-km binary objects
considered here may have been quite common as the very last portions of the Solar System's
gas disk were being depleted.
Finally, the known irregular satellite population numbers < 100, and many of these are
probably members of families – thus, we need only produce at most a few dozen captures.
While accounting for the origin of such a small population is difficult, our simulations show
that it is likely binaries played a role. We conclude by offering our model as a new idea that
alleviates many but not all of the problems faced by previous models, but acknowledge that
the without a detailed understanding of the initial population including binary statistics, a
firm conclusion is not possible.
– 24 –
REFERENCES
Agnor, C. B. and Hamilton, D. P., 2006a. Neptune's capture of its moon Triton in a binary-
planet gravitational encounter. Nature 441, 192-194.
Agnor, C. B. and Hamilton, D. P., 2006b. Satellite capture via binary-planet gravitational
encounters. Bull. Am. Astron. Soc. 38, 674.
Astakhov, S. A. and Farrelly, D., 2004. Capture and escape in the elliptic restricted three-
body problem. Mon. Not. R. Astron. Soc. 354, 971-979.
Astakhov, S. A., Burbanks, A. D., Wiggins, S., and Farrelly, D., 2003. Chaos-assisted capture
of irregular moons. Nature 423, 264-267
Bate, M. R., Lubow, S. H., Ogilvie, G. I., and Miller, K. A., 2003. Three-dimensional
calculations of high- and low-mass planets embedded in protoplanetary discs. Mon.
Not. R. Astron. Soc. 341, 213-229.
Canup, R. M. and Ward, W. R., 2002. Formation of the Galilean satellites: Conditions of
accretion. Astron. J. 124, 3404-3423.
Cieza, L. and 19 colleagues, 2007. The Spitzer c2d survey of weak-line T Tauri stars. II. New
constraints on the timescale for planet building. Astrophys. J. 667, 308-328.
Colombo, G. and Franklin, F. A., 1971. On the formation of the outer satellite groups of
Jupiter. Icarus 15, 186-189.
´Cuk, M. and Burns, J. A., 2004a. Gas-drag-assisted capture of Himalia's family. Icarus 167,
369-381.
´Cuk, M. and Burns, J. A., 2004b. On the secular behavior of irregular satellites. Astron. J.
128, 2518-2541.
Cuzzi, J. N., Hogan, R. C., and Shariff, K., 2008. Toward planetesimals: Dense chondrule
clumps in the protoplanetary nebula. Astrophys. J. 687, 1432-1447.
D'Angelo, G., Henning, T., and Kley, W., 2003. Thermohydrodynamics of circumstellar disks
with high-mass planets. Astrophys. J. 599, 548-576.
Hamilton, D. P. and Burns, J. A., 1991. Orbital stability zones about asteroids. Icarus 92,
118-131.
Henon, M., 1969. Numerical exploration of the resticted problem, V. Astron. Astrophys. 1,
223-238.
– 25 –
Heppenheimer, T. A. and Porco, C., 1977. New contributions to the problem of capture.
Icarus 30, 385-401.
Jewitt, D. and Sheppard, S., 2005. Irregular satellites in the context of planet formation.
Space Science Reviews 116, 441-455.
Johansen, A., Oishi, J. S., Low, M.-M. M., Klahr, H., Henning, T., and Youdin, A., 2007.
Rapid planetesimal formation in turbulent circumstellar disks. Nature 448, 1022-1025.
Lubow, S. H., Seibert, M., and Artymowicz, P., 1999. Disk accretion onto high-mass planets.
Astrophys. J. 526, 1001-1012.
Merline, W. J. and 11 colleagues, 2007. The search for Trojan binaries. Bull. Am. Astron.
Soc. 38, 538.
Morbidelli, A., Bottke, W. F., Nesvorn´y, D., and Levison, H. F., 2009. Asteroids were born
big. Icarus, in press.
Mosqueira, I. and Estrada, P. R., 2003. Formation of the regular satellites of giant planets
in an extended gaseous nebula I: Subnebula model and accretion of satellites. Icarus
163, 198-231.
Murray, C. D. and Dermott, S. F., 1999. Solar System Dynamics. Cambridge Univ. Press,
Cambridge, UK.
Nesvorn´y, D., 2008. Formation of Kuiper belt binaries. Bull. Am. Astron. Soc. 40, 464.
Nesvorn´y, D., Alvarellos, J. L. A., Dones, L., and Levison, H. F., 2003. Orbital and collisional
evolution of the irregular satellites. Astron. J. 126, 398-429.
Nesvorn´y, D., Vokrouhlick´y, D., and Morbidelli, A., 2007. Capture of irregular satellites
during planetary encounters. Astron. J. 133, 1962-1976.
Noll, K. S., Grundy, W. M., Stephens, D. C., Levison, H. F., Kern, S. D., 2008. Evidence
for two populations of classical transneptunian objects: The strong inclination de-
pendence of classical binaries. Icarus 194, 758-768.
Papaloizou, J. C. B., Nelson, R. P., Kley, W., Masset, F. S., and Artymowicz, P., 2007.
Disk-planet interactions during planet formation. In: Reipurth, D., Jewitt, D., and
Keil, K. (Eds.), Protostars and Planets V, Univ. of Arizona Press, Tucson, 655-668.
Pollack, J. B., Burns, J. A., and Tauber, M. E., 1979. Gas drag in primordial circumplanetary
envelopes – a mechanism for satellite capture. Icarus 37, 587-611.
– 26 –
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak, M., and Greenzweig,
Y., 1996. Formation of the giant planets by concurrent accretion of solids and gas.
Icarus 124, 62-85.
Rauch, K. P. and Hamilton, D. P., 2002. The HNBody package for symplectic integration of
nearly-Keplerian systems. Bull. Am. Astron. Soc. 34, 938.
Saha, P. and Tremaine, S., 1993. The orbits of the retrograde Jovian satellites. Icarus 106,
549-562.
Sheppard, S. S. and Jewitt, D. C., 2003. An abundant population of small irregular satellites
around Jupiter. Nature 423, 261-263.
Silverstone, M. D. and 16 colleagues, 2006. Formation and Evolution of Planetary Systems
(FEPS): Primordial warm dust evolution from 3 to 30 Myr around Sun-like stars.
Astrophys. J. 639, 1138-1146.
Strutskie, M. F., Dutkevitch, D., Strom, S. E., Edwards, S., Strom, K. M., and Shure,
M. A., 1990. A sensitive 10-micron search for emission arising from circumstellar dust
associated with solar-type pre-main-sequence stars. Astron. J. 99, 1187-1195.
Tsiganis, K., Gomes, R., Morbidelli, A., and Levison, H. F., 2005. Origin of the orbital
architecture of the giant planets of the Solar System. Nature 435, 459-461.
Whipple, A. L. and Shelus, P. J., 1993. A secular resonance between Jupiter and its eighth
satellite? Icarus 101, 265-271.
Vokrouhlick´y, D., Nesvorn´y, D., and Levison, H. F., 2008. Irregular satellite capture by
exchange reactions. Astron. J. 136, 1463-1476.
This preprint was prepared with the AAS LATEX macros v5.2.
– 27 –
Table 1. Mass Ratio Tests For Binaries With CJ =3.03
Total Binary
Mass Component
Ratio Radii (km) Mass (1019kg)
Separation m1 Capture m2 Capture Total 1C Capture
(km)
Percentage
Percentage
Percentage
3:2
1:1
1:1
4:1
1:1
1:1
113-98
85-85
127-127
124-78
42-42
169-169
2.0
1.0
3.4
2.0
0.13
8.1
1350
1080
1620
1160
465
1860
0.54
0.58
0.94
0.21
0.24
1.29
0.96
0.58
0.94
1.17
0.24
1.29
1.50
1.16
1.88
1.38
0.48
2.58
Note. - Two experiments with binaries of unequal mass. The first four columns identify properties
of the binary, while the final three columns list capture statistics (with m1 ≥ m2). We used the impulse
approximation (Eqs. 1 and 2) to determine the properties of equivalent equal-mass binaries that match
the tidal disruption radius and speed of either m1 (boldfaced) or m2 (italicized).
In all cases, our
predicted percentages agree with the actual measurements to within about 10%.
– 28 –
Fig. 1.- The speeds of unequal-mass binary components (v1 and v2) and the center-of-mass
speed (vCM ) relative to Jupiter, where m1 > m2. Since the binary center of mass approaches
the planet along a hyperbolic trajectory, it is always traveling faster than the escape speed
(vesc). In this example, the smaller component's speed dips below the escape speed for a
portion of its orbit. If the binary is disrupted during this interval, the smaller component
will be captured by the planet.
– 29 –
3 . 0 1
3 . 0 1 5
3.0 2
3 . 0 2 5
3
.
0
3
7
3
3 . 0
Hill sphere
∗
•J
3.035
∗
3.02
3
.
0
3
5
3.01
1
x (aJ)
1.05
9
3
0
.
3
1.1
0.15
0.1
0.05
)
J
a
(
y
0
−0.05
−0.1
3.0
3
9
9
3
3.0
7
3
3.0
3
3.0
5
2
3 . 0
3.015
0.85
0.9
0.95
Fig. 2.- The Hill sphere of Jupiter (central dot marked 'J') with zero-velocity curves (ZVCs)
corresponding to the labeled Jacobi constant (CJ ) values. The Sun is to the left at (0,0) and
the Jupiter-Sun separation (aJ ) is used as the unit of distance. The two asterisks are located
at the planet's L1 and L2 Lagrange points. A body near Jupiter with CJ & 3.0387 (e.g.,
the curve for CJ = 3.039) has ZVCs enclosed around the planet, and it would be unable to
escape.
– 30 –
15
10
5
0
1.04
1.02
1
0.98
0.96
3.04
3.038
3.036
3.034
0
2
4
6
8
10
12
Fig. 3.- Distance from the binary components (each 65-km in radius) to Jupiter (r) is
plotted in units of the tidal disruption radius (rtd ≈ 70 RJ ) in the top panel; the middle
panel shows the binary's separation (rB) over time, in units of its initial separation (2aB);
and the bottom panel displays the Jacobi constant of each component. One component is
plotted with a solid line and the other with a dotted line throughout the figure. The dashed
line in the top panel indicates the tidal disruption radius. In the bottom panel, the dashed
line represents the critical Jacobi constant.
– 31 –
Fig. 4.- The orbits of the binary components discussed in Fig. 3. Jupiter is the dot near the
center of each panel, and the Sun is located to the left at (0,0). The components are plotted
together in the top panel, with solid/dotted lines corresponding to those in Fig. 3. The
objects are bound to each other as they approach Jupiter from the left side of each panel. A
close approach to the planet tidally disrupts the binary, sending one component (dotted line)
out of the system and causing the other component (solid line) to be permanently captured
by the planet. The bottom panel contains only the orbit of the component that escapes,
with an arrow marking the location where the binary's separation first exceeds 102% of its
original value. About equidistant on the other side of the planet, the binary's separation is
more than twice its initial value.
– 32 –
Fig. 5.- Inclination of 100-km captured binary components as a function of initial Jacobi
constant, plotted for the bodies' first close approaches. Similar plots show that the relation-
ship holds for all close approaches of binaries or single bodies, captured or not. Also plotted
is a line connecting the mean at each Jacobi constant (marked by the stars) as well as 1-σ
error bars. See Section 4.1 for discussion on the CJ-inclination relationship.
– 33 –
1
0.1
0.01
10
100
1000
Fig. 6.- Three modes of capture for integrations of RB = 65-km binary pairs with CJ =
3.037, as a function of the separation of the binary (rB): one component captures ('1C'),
the binary splits and both capture independently ('2C-IND'), or the binary remains bound
and captures as a pair ('2C-BIN'). We plot the capture percentage for objects (rather than
binaries) to facilitate comparison with single bodies (upper solid line). Recall our working
definition of capture to mean bodies still orbiting the planet after 1,000 years. Note that the
2C-BIN curve approaches the value for singles at small separations while the sum of the 1C
and 2C-IND captures rates approaches the same value for large separations.
– 34 –
Fig. 7.- Capture percentages for binaries of different masses compared to single objects.
All bodies were started on Jupiter's Hill sphere. For each binary mass, a single separation
was used over the range of initial Jacobi constants: 471 km = 7.25 binary radii (RB) for
the 65-km binaries, 1225 km = 12.25 RB for the 100-km set, and 1512 km = 12.10 RB for
125-km binaries. These separations give near maximum capture rates for the majority of the
Jacobi constants tested, with the exception of the highest CJ, where the optimum separation
is closer to 20 binary radii.
– 35 –
Fig. 8.- Single objects integrated from two launch distances. True captures curves (with
points that are solid circles) show only those captures that originated far from Jupiter, while
curves that show true plus contaminated captures (with open triangles) also include objects
that were stably orbiting the planet at the beginning of the integrations. For launch on the
Hill sphere (1.0 rH ), we see that the prograde orbits (high CJ) are not affected at all, while
many of the retrograde orbits (low CJ ) are discovered to be false. For launches at 1.1 rH,
the true and true-plus-contaminated curves overlap completely, showing that contamination
by false captures is effectively zero.
– 36 –
3
2
1
0
3.01
3.02
3.03
Fig. 9.- Capture rates of 100-km binaries that were launched from 1.1 rH (upper dashed
line), 1.2 rH (solid line), and expected capture rates for 1.2 rH (dotted line) calculated by
scaling the 1.1 rH rates by the percent of 1.2 rH trajectories crossing interior to the 1.1 Hill
radii. All are shown with 1-σ error bars. This scaling equalizes the capture percentages
between the two launch distances to within 20%. For reference, the capture rate for single
bodies starting from 1.1 rH is also plotted; compare with Fig. 7.
– 37 –
Retrograde
Prograde
Retrograde
Prograde
0.8
0.6
0.4
0.2
0
1C
Single
2C-BIN
0.8
0.6
0.4
0.2
0
1C
Single
2C-BIN
3.01
3.02
3.03
3.01
3.02
3.03
Jacobi constant
Jacobi constant
Fig. 10.- Modes of capture vs. Jacobi constant, for integrations of 100-km binaries starting
at 1.0 rH (left panel) and 1.1 rH (right panel). The 1C curve peaks are off the top of the
plot at ∼4.2% for the 1.0 rH runs and ∼3.0% for the 1.1 rH group. The 2C-IND capture
rates are extremely small for both starting distances (e.g., Fig. 6) and, for clarity, are not
plotted here. The capture rates of single bodies are plotted for comparison.
– 38 –
Fig. 11.- The percent of captured objects that do not cross interior to 26 RJ (Callisto's
semi-major axis) during the 1,000-year integrations. The bodies were started on the Hill
sphere. Compare with the entire set of captures seen in Fig. 7.
– 39 –
Fig. 12.- An example of a simple gas drag applied to a prograde (left panels; the same
initial orbit as in Figs. 3 and 4) and a retrograde orbit (right panels). The orbits are shown
immediately after capture in the upper panels, and the bottom panels show the orbits after
25,000 years of evolution. Jupiter is the dot in the center, and the Sun is to the left at (0,0).
These orbits are the result of 1C capture from equal-mass binaries with 65-km components.
Before disruption, the prograde binary had a separation of 70 RB and a Jacobi constant of
∼ 3.037. The retrograde binary had an initial CJ of ∼ 3.003 and was initially separated by
460 RB.
|
1907.08187 | 1 | 1907 | 2019-07-18T17:57:52 | Martian Year 34 Column Dust Climatology from Mars Climate Sounder Observations: Reconstructed Maps and Model Simulations | [
"astro-ph.EP",
"physics.ao-ph"
] | We have reconstructed longitude-latitude maps of column dust optical depth (CDOD) for Martian year (MY) 34 (May 5, 2017 --- March 23, 2019) using observations by the Mars Climate Sounder (MCS) aboard NASA's Mars Reconnaissance Orbiter spacecraft. Our methodology works by gridding standard and newly available estimates of CDOD from MCS limb observations, using the "iterative weighted binning" methodology. In this work, we reconstruct four gridded CDOD maps per sol, at different Mars Universal Times. Together with the seasonal and day-to-day variability, the use of several maps per sol allows to explore also the daily variability of CDOD in the MCS dataset, which is shown to be particularly strong during the MY 34 equinoctial Global Dust Event (GDE). Regular maps of CDOD are then produced by daily averaging and spatially interpolating the irregularly gridded maps using a standard "kriging" interpolator, and can be used as "dust scenario" for numerical model simulations. In order to understand whether the daily variability of CDOD has a physical explanation, we have carried out numerical simulations with the "Laboratoire de M\'et\'eorologie Dynamique" Mars Global Climate Model. Using a "free dust" run initiated at $L_s \sim 210^\circ$ with the corresponding kriged map, but subsequently free of further CDOD forcing, we show that the model is able to account for some of the observed daily variability in CDOD. The model serves also to confirm that the use of the MY 34 daily-averaged dust scenario in a GCM produces results consistent with those obtained for the MY 25 GDE. | astro-ph.EP | astro-ph |
MARTIAN YEAR 34 COLUMN DUST CLIMATOLOGY FROM
MARS CLIMATE SOUNDER OBSERVATIONS: RECONSTRUCTED
MAPS AND MODEL SIMULATIONS
A PREPRINT
Luca Montabone
Space Science Institute, Boulder, CO, USA
Laboratoire de Météorologie Dynamique (LMD/IPSL), Sorbonne Université,
Centre National de la Recherche Scientifique, École Polytechnique,
École Normale Supérieure, Paris, France
Laboratoire de Météorologie Dynamique (LMD/IPSL), Sorbonne Université,
Aymeric Spiga
Centre National de la Recherche Scientifique, École Polytechnique,
École Normale Supérieure, Paris, France
and
and
Institut Universitaire de France, Paris, France
Jet Propulsion Laboratory, Caltech Institute of Technology, Pasadena, CA, USA
David M. Kass and Armin Kleinböhl
Laboratoire de Météorologie Dynamique (LMD/IPSL), Sorbonne Université,
François Forget and Ehouarn Millour
Centre National de la Recherche Scientifique, École Polytechnique,
École Normale Supérieure, Paris, France
Corresponding author: Luca Montabone ([email protected])
July 19, 2019
ABSTRACT
We have reconstructed longitude-latitude maps of column dust optical depth (CDOD) for Martian
year (MY) 34 (May 5, 2017 -- March 23, 2019) using observations by the Mars Climate Sounder
(MCS) aboard NASA's Mars Reconnaissance Orbiter spacecraft. Our methodology works by gridding
standard and newly available estimates of CDOD from MCS limb observations, using the "iterative
weighted binning" methodology. In this work, we reconstruct four gridded CDOD maps per sol, at
different Mars Universal Times. Together with the seasonal and day-to-day variability, the use of
several maps per sol allows to explore also the daily variability of CDOD in the MCS dataset, which is
shown to be particularly strong during the MY 34 equinoctial Global Dust Event (GDE). Regular maps
of CDOD are then produced by daily averaging and spatially interpolating the irregularly gridded
maps using a standard "kriging" interpolator, and can be used as "dust scenario" for numerical model
simulations. In order to understand whether the daily variability of CDOD has a physical explanation,
we have carried out numerical simulations with the "Laboratoire de Météorologie Dynamique" Mars
Global Climate Model. Using a "free dust" run initiated at Ls ∼ 210◦ with the corresponding kriged
map, but subsequently free of further CDOD forcing, we show that the model is able to account for
A PREPRINT - JULY 19, 2019
some of the observed daily variability in CDOD. The model serves also to confirm that the use of the
MY 34 daily-averaged dust scenario in a GCM produces results consistent with those obtained for
the MY 25 GDE.
Plain Language Summary
Large dust storms on Mars have dramatic impacts on the entire atmosphere, but may also have critical consequences
for robotic and future human missions. Therefore, there is compelling need to produce an accurate reconstruction of
their spatial and temporal evolution for a variety of applications, including to guide Mars climate model simulations.
The recently ended Martian year 34 (May 5, 2017 -- March 23, 2019) represents a very interesting case because an
extreme dust event occurred near the time of the northern autumn equinox, consisting of multiple large dust storms
engulfing all longitudes and most latitudes with dust for more than 150 Martian days ("sols"). We have used satellite
observations from the Mars Climate Sounder instrument aboard NASA's Mars Reconnaissance Orbiter to reconstruct
longitude-latitude maps of the opacity of the atmospheric column due to the presence of dust at several times in each sol
of Martian year 34. These maps allow us to analyze the seasonal, day-do-day, and day-night variability of dust in the
atmospheric column, which is particularly intense during the extreme dust event. We have also used simulations with a
Mars climate model to show that the strong day-night variability may be partly explained by the large-scale circulation.
1
Introduction
Martian dust aerosols are radiatively active, and the dust cycle -- lifting, transport, and deposition -- is considered to be
the key process controlling the variability of the Martian atmospheric circulations on a wide range of time scales (see e.g.
the recent review by Kahre et al., 2017, and references therein). Dust storms are the most remarkable manifestations
of this cycle, and one of the most crucial weather phenomena in need to be studied to fully understand the Martian
atmosphere.
Martian dust storms are: 1. a source of strong atmospheric radiative forcing, and alteration of surface energy budget (e.g.
Streeter et al., 2019); 2. a major component of the atmospheric inter-annual, seasonal, day-to-day, and daily variability
[e.g. Kleinböhl et al., this issue]; 3. a way to redistribute dust on the planet via long-range particle transport; 4. a key
element to also understand longer-term surface changes, such as, for example, the aeolian erosion, and the retreat of the
seasonal ice cap; 5. a strong disturbance of temperature and density propagating from the lower to the upper atmosphere,
including the lower thermosphere, the ionosphere, and the magnetosphere [e.g. Girazian et al., and Xiaohua et al., this
issue]; 6. a cause of increase of loss of chemical species via escape (Fedorova et al., 2018; Heavens et al., 2018, and
Xiaohua et al., this issue); 7. a source of hazards for spacecraft Entry, Descending and Landing (EDL) manoeuvres, for
operations by solar-powered surface assets, and for future robotic and human exploration (e.g. Levine et al., 2018).
Dust storms on Mars can be studied by using a variety of approaches: analysis of observations from satellites and
landers/rovers, numerical simulations from Global Climate Models (GCM), and data assimilation techniques.
One of the most dramatic and (so far) unpredictable event linked to Martian dust storms is the onset of a Global Dust
Event -- hereinafter GDE. In the literature, these events are also named "Planet-Encircling Dust Storms" (e.g. Cantor,
2007), or "Global Dust Storms". Here we choose the terminology "Global Dust Event" (already discussed and used in
Montabone and Forget, 2018) because 1. even large regional storms can inject dust high enough in the atmosphere,
which eventually encircles the planet, and 2. these dust events are usually characterized by several storms occurring
simultaneously, or one after the other one in rapid succession.
In the last Martian decade (Martian Year -- hereinafter MY -- 25 to 34), spanning nearly two Earth decades from 2000
to 2019, three GDEs occurred: an equinoctial event in MY 25, a solstitial event in MY 28, and another equinoctial event
in MY 34, starting only few mean solar days -- sols -- after the corresponding onset of the MY 25 event. GDEs inject
a large amount of dust particles in the Martian atmosphere, strongly modify the thermal structure and the atmospheric
dynamics during several months (i.e. several tens degrees of solar longitude -- Ls -- , see e.g. Wilson and Hamilton,
1996; Montabone et al., 2005), and impact the Martian water cycle and escape rate (Fedorova et al., 2018; Heavens
et al., 2018). Similar events were previously observed in Martian Years 1, 9, 10, 12, 15, and 21 (Martin and Zurek,
1993; Cantor, 2007; Montabone and Forget, 2018; Sánchez-Lavega et al., 2019). The inter-annual variability of GDEs
is irregular, and likely controlled at the first-order by the redistribution of dust on Mars over the timescale of a few years
(Mulholland et al., 2013; Vincendon et al., 2015).
The latest equinoctial GDE had its initial explosive growth in early northern fall of MY 34 (Ls approximately in the
range 185 − 190◦, i.e. late May 2018 -- early June 2018). A regional dust storm started near the location of the Mars
Exploration Rover "Opportunity". The visible opacity quickly reached a very high value of 10.8, which led to the
end-of-mission of the Opportunity rover, with last communication received on June 10, 2018. The regional dust storm
2
A PREPRINT - JULY 19, 2019
then moved southward along the Acidalia storm-track, and expanded both in the northern hemisphere from eastern
Tharsis to Elysium (including the location of the Mars Science Laboratory "Curiosity" rover and the landing site of
InSight), and towards the southern hemisphere (Malin et al., 2018).
The evolution of this MY 34 GDE in summer 2018 has been closely monitored by three NASA's orbiters, including the
Mars Reconnaissance Orbiter (MRO) and its Mars Climate Sounder (MCS) instrument [Kass et al. this issue], two ESA's
orbiters (Mars Express and ExoMars Trace Gas Orbiter), the ISRO's Mangalyaan orbiter, and ground-based telescopes
(Sánchez-Lavega et al., 2019). It has also been observed in details from the surface by the Curiosity rover, which could
still operate in dust-storm conditions thanks to its nuclear-powered system. From meteorological observations carried
out aboard Curiosity with the Rover Environmental Monitoring Station, Guzewich et al. (2018) concluded that the local
optical depth reached 8.5, the incident total UV solar radiation at the surface decreased by 97%, the diurnal range of
air temperature decreased by 30 K, and the semidiurnal pressure tide amplitude increased to 40 Pa. Curiosity did not
witness dust being lifted within the Gale Crater site, which indicates that the increase in dust loading at its location
corresponds to dust particles transported in the Gale Crater area by large-scale and regional winds.
Beyond the undoubtedly interesting GDE, MY 34 also features the development of an unusual late-winter large regional
storm whose peak CDOD is only rivaled by that in MY 26, reaching 75% of its peak value. It is, however, reminiscent
of the two global events that were successively monitored by the Viking landers in 1977 at Ls = 205◦ and Ls = 275◦
(Ryan and Henry, 1979; Zurek, 1982). Overall, therefore, MY 34 represents a unique year for studies linked to the
onset/evolution of dust storms, and their impact on the entire Martian atmospheric system. Consequently, there is a
compelling need to produce an accurate reconstruction of the spatial and temporal evolution of the dust optical depth
in MY 34, particularly covering the GDE, but also putting the unprecedented weather measurements acquired by the
InSight lander during the late-winter regional storm in the global context (Spiga et al., 2018).
Montabone et al. (2015) have developed a methodology to grid values of column dust optical depth (CDOD) retrieved
from multiple polar orbiting satellite observations, such as NASA's Mars Global Surveyor (MGS), Mars Odyssey
(ODY), and MRO. Using this methodology (a combination of "Iterative Weighted Binning" -- IWB -- and kriging
spatial interpolation), they were able to produce a multi-annual dataset of daily CDOD maps extending from MY 24
to MY 33, which is publicly available on the Mars Climate Database (MCD) project webpage at http://www-mars.
lmd.jussieu.fr/ (see "Martian dust climatology").
In this paper we describe how we made use of newly processed dust opacity retrievals from thermal infrared observations
of the MRO/MCS instrument (McCleese et al., 2007) in order to reconstruct maps of column dust optical depth
specifically for MY 34, and describe aspects of the two-dimensional dust climatology. In Section 2, we discuss the
improvements both to the MCS retrievals and to the gridding methodology described in Montabone et al. (2015). In
Section 3 we analyze in general terms the CDOD variability at seasonal scale, and specifically the day-to-day evolution
of the GDE and late-winter dust storm. We also address the daily variability observed when reconstructing multiple
CDOD maps per sol. In Section 4, the generated "MY 34 dust scenario" (i.e. complete, daily, kriged maps) is used in
simulations with the Laboratoire de Météorologie Dynamique Mars GCM (LMD-MGCM) to 1. assess that the impact
of the MY 34 GDE on the Martian atmosphere is well accounted for, and 2. verify that the GCM is able to produce
some of the diurnal variability observed in the reconstructed multiple CDOD maps per sol, when the dust scenario is
only used as initial condition. Conclusions are drawn in Section 5. Appendix A is devoted to clarify the versioning of
the CDOD maps.
2 Building column dust optical depth maps
The methodology described in Montabone et al. (2015) to grid CDOD values using the IWB, and to spatially interpolate
the daily maps using kriging, has been applied to observations by MGS/Thermal Emission Spectrometer (TES),
ODY/Thermal Emission Imaging System (THEMIS), and MRO/MCS from MY 24 to MY 32. For MY 33, because of
the progressive change in local time of THEMIS observations, the weighting functions of the THEMIS dataset and the
MCS dataset were slightly modified to favor the MCS dataset. As is mentioned in the introduction, version 2.0 (v2.1 for
MY 33) of both irregularly gridded maps and regularly kriged ones ("dust scenarios") are available on the MCD project
website.
For the specific case of the MY 34 GDE, the MCS team has updated their retrievals of temperature, dust and water ice
profiles. We have correspondingly updated the gridding/kriging methodology with the aim of producing a more refined
and accurate climatology, both for scientific studies and for the use in numerical model simulations. Therefore, in the
following we describe how we reconstruct CDOD maps for MY 34 (currently version 2.5). We provide some details
about the differences between current and previous versions (i.e. v2.2, v2.3, and v2.4) in Appendix A.
3
A PREPRINT - JULY 19, 2019
2.1 Observational dataset
In MY 34, single THEMIS CDOD retrievals are no longer available. Because of the late local time of THEMIS
observations in MY 34, Smith et al. [this issue] had to develop a "stacking" algorithm that assesses how a group of
THEMIS spectra in a solar longitude/latitude bin change as a function of estimated thermal contrast. Therefore, we do
not use THEMIS any more in MY 34 and we completely rely on estimated CDODs from MCS.
Dust opacity retrievals from thermal infrared observations of the Mars Climate Sounder instrument aboard MRO are
described in Kleinböhl et al. (2009); Kleinböhl et al. (2011); Kleinböhl et al. (2017a). The currently standard MCS
dataset, based on the v5.2 "two-dimensional" retrieval algorithm specifically described in (Kleinböhl et al., 2017a), has
been reprocessed by the MCS team during the time of the MY 34 GDE to obtain better coverage in the vertical and,
therefore, more reliable estimates of CDOD values during the event [Kleinböehl et al., this issue]. This latest MCS
dataset, only available between May 21, 2018 (Ls ∼ 179◦) and October 15, 2018 (Ls ∼ 269◦), labelled v5.3.2, is an
interim version that includes the use of a far infrared channel for retrievals of dust. The differences between MCS
retrievals version 5.2 and 5.3 are as follows:
• Use of B1 detectors to extend the dust profile retrieval: the dust extinction efficiency in channel B1 at 32 µm is
only about half the value of channel A5 at 22 µm (Kleinböhl et al., 2017b), which is the primary channel for
dust retrievals, allowing profiles to extend deeper by 1 to 1.5 scale heights;
• Accepting more opacity from aerosols in the temperature retrieval channel A3;
• Modifications for determining surface temperature when there are no matching on-planet views (primarily
cross-track views) to improve the performance under high dust conditions when the array is lifted and limb
views do not intersect the surface.
CDODs are estimated by integrating the dust opacity profiles after an extrapolation from the lowest altitude at which
profile information is available, under the assumption of homogeneously mixed dust [see Fig. 1 of Kleinböehl et al.,
this issue]. For the reconstructed CDOD maps in MY 34, we use MCS v5.3.2 estimated CDODs from Ls ∼ 179◦ to
Ls ∼ 269◦, and MCS v5.2 otherwise.
2.2 Data Quality Control
A general discussion about the limitations of using CDOD estimates from MCS is included in Montabone et al. (2015),
specifically Section 2.1.2. As is mentioned in the previous sub-section, the extended vertical coverage in MCS v5.3.2
helps estimate CDODs more accurately. In the present work, therefore, we have improved the definition of the Quality
Control (QC) procedure with respect to the one used in Montabone et al. (2015), particularly by allowing a more
extensive use of dayside observations. We define dayside observations as those with local times in the range 09:00 < lt
≤ 21:00, although most of dayside observations at low latitudes have local times close to 15:00. Nightside observations
are defined as those outside the dayside range, with most nightside observations at low altitude having local times close
to 03:00. Note that MCS is also able to observe cross-track, thus providing information in a range of local times at
selected positions during the MRO orbits (Kleinböhl et al., 2013). We have also better defined a dust quality flag in
MCS v5.3.2 to help filtering those observations where a significant number of detectors were excluded in the retrieval
of the dust opacity profile, because of radiance residuals exceeding threshold values (Kleinböhl et al., 2009). Each
excluded detector corresponds to a truncation of ∼ 5 km from the altitude range that was originally selected by the
retrieval algorithm based on line-of-sight opacity.
We apply the following QC procedure to the MCS CDOD values at 21.6 µm in extinction:
• To discard values when they are most likely contaminated by CO2 ice (i.e. if, at any level below 40 km altitude,
• To discard values when water ice opacity is greater than dust opacity at the cut-off altitude of the corresponding
the temperature is T < TCO2 + 10K, and the presumed dust opacity is larger than 10−5 km−1);
dust profile;
• To discard cross-track CDODs with cut-off altitudes higher than 8 km (i.e. the corresponding dust opacity
profiles do not extend down to 8 km altitude or lower);
• To discard dayside values with cut-off altitudes higher than 8 km unless it is MCS v5.3.2 data or MCS v5.2
data during the late-winter dust storm (i.e. for Ls ≤ 179◦, 269◦ < Ls ≤ 312◦, and Ls ≥ 350◦). Additionally,
we discard dayside CDODs with cut-off altitudes higher than 16 km in a short period before the onset of the
GDE (179◦ < Ls < 186.5◦) and during the late-winter dust storm (312◦ < Ls < 350◦). No filtering based
on cut-off altitude is applied during the GDE for dayside CDODs;
4
A PREPRINT - JULY 19, 2019
• To discard CDODs when more than 1 detector is excluded inside the limits of the MCS v5.3.2 (179◦ ≤ Ls ≤
269◦) as well as if any detector is excluded in MCS v5.2 during the late-winter dust storm (312◦ < Ls < 350◦);
• To assign a fixed value of 0.01 to very low values of CDOD < 0.01 having cut-off altitude higher than 4 km.
We plot in Fig.1 the percentage number of CDOD values that are flagged by each individual filter, together with the
total of the filtered values after the application of the complete QC procedure. The total does not correspond to the sum
of each single filter, as a CDOD value can be flagged by multiple filters. This figure clearly shows that the presence
of water ice in spring and summer strongly affects the number of CDOD values passing the QC. Dayside values are
also problematic because their corresponding dust profiles usually have rather high cut-off altitudes, compared to
nightside values. Cross-track values have the tendency to exhibit rather high cut-off altitudes as well, and lead to
questionable column dust optical depths. As a consequence, a large number of them at low- and mid-latitudes are
discarded throughout the year. Observations where more than one detector was excluded in the retrieval are about
20% throughout MCS v5.3.2. The number of values filtered because of possible carbon dioxide ice contamination is
relatively low throughout the year (less than 10%).
Figure 1: Percentage number of CDOD values flagged by each individual filter in the QC procedure within 30◦
solar longitude ranges in MY 34 (color lines), together with the percentage total number of filtered CDODs after the
application of the complete QC procedure (black line). The numbers are associated to the middle of each 30◦ solar
longitude range. Note that the "excluded detector" filter does not apply before Ls = 179◦ and for 269◦ < Ls < 312◦,
and that the "dayside" filter does not apply during the GDE (186.5◦ ≤ Ls ≤ 269◦).
Note that, despite the improvements in MCS v5.3.2, the main issue for estimating CDODs using limb observations
remains the fact that many dust opacity profiles have rather high cut-off altitudes (particularly dayside ones, see also
third column of Fig.3), because of either large dust or water ice opacities. This implies to extrapolate the profiles for a
long altitude range under the assumption of homogeneously mixed dust, thus increasing the uncertainty on the column
values.
After QC, the number of available values is plotted in Fig.2, separated between nightside and dayside values. The
aphelion cloud belt and the winter polar hoods are mainly responsible for the lack of data at equatorial latitudes in the
dayside plot, and at high latitudes in both dayside and nightside plots. The implementation of the new "water ice" filter
is effective in reducing the probability that the lowest levels of the dust profiles are contaminated by the presence of
clouds, but there is a risk of filtering out retrievals that actually may have been usable, particularly at high latitudes. A
refinement of this filter can be addressed in future work. Vertical bands with no data are periods when MCS did not
observe.
2.3 Data uncertainties and processing
Together with the QC procedure, we have also revised the empirical method to estimate the uncertainties on the MCS
CDOD values at 21.6 µm in extinction, with respect to the one used in Montabone et al. (2015). We apply the following
relative uncertainties:
5
A PREPRINT - JULY 19, 2019
Figure 2: Number of nightside (upper panel) and dayside (lower panel) values of column dust optical depth available
for gridding after passing the quality control procedure described in the text. The number of values is summed in 1
sol × 2◦ latitude bins, and plotted as a function of sol from the beginning of MY 34, corresponding solar longitude,
and latitude. Dayside observations are defined to have local times in the range ]09:00, 21:00], whereas nightside
observations are defined to have local times in [00:00, 09:00] and ]21:00, 24:00].
• 10% for CDOD values < 0.01 having cut-off altitude higher than 4 km (i.e. for those values replaced with
CDOD=0.01);
• 5% for CDOD values < 0.01 or values with cut-off altitudes lower than 4 km;
• When CDOD ≥ 0.01 or cut-off altitude ≥ 4 km, we assign the largest relative uncertainty between the one
calculated as a linear function of CDOD and the one calculated as a linear function of the cut-off altitude. The
two functions are defined in such a way that the uncertainty is 5% if CDOD = 0.01 or cut-off altitude = 4 km,
20% if CDOD = 1.5, and 30% if cut-off altitude = 25 km.
As detailed in Montabone et al. (2015), further data processing consists in converting MCS CDODs from 21.6 µm in
extinction to absorption-only 9.3 µm by multiplying by 2.7, to be consistent with the climatologies of the previous
Martian years. We then normalize the values to the reference pressure level of 610 Pa, but instead of using the surface
pressure value calculated by the MCD pres0 routine (Forget et al., 2007), we now use the same surface pressure value
used for the corresponding MCS retrieval. MCS retrieves pressure at the pointing altitude where it is most sensitive
to pressure (typically 20-30 km (Kleinböhl et al., 2009)), from which surface pressure can be extrapolated with an
uncertainty estimate based on pointing uncertainty. In conditions where a pressure retrieval is unsuccessful (typically in
conditions of high aerosol loading) the MCS algorithm uses pressure derived from the climatological Viking surface
pressure (Withers, 2012). In this case, the uncertainty of the surface pressure is derived from the day-to-day root mean
squared of surface pressure from the MCD v5.3, interpolated at the specific location and season of an observation using
a pre-built 5◦ solar longitude × 5◦ latitude array (as described in Section 2.3 of Montabone et al., 2015).
2.4 Gridding methodology
In this work we closely follow the basic principles of reconstructing CDOD maps, which have been detailed in
Montabone et al. (2015). Iterative Weighted Binning (IWB) is applied to CDOD values at 9.3 µm in absorption,
normalized to 610 Pa, together with their corresponding uncertainties, to produce gridded values on a 6◦ longitude × 5◦
latitude map, with possible "no value" assigned at locations where data do not satisfy given requirements. The current
6
A PREPRINT - JULY 19, 2019
criterion to accept a value of weighted average at a particular grid point at any given iteration is that there must be at
least one observation within a distance of 200 km from the grid point, while the other parameters listed for MCS in
Table 1 of Montabone et al. (2015) remain the same.
The key difference with respect to the methodology described in Section 3 of Montabone et al. (2015) is that in this work
we create four gridded maps per sol at four different Mars Universal Times (MUT, i.e. the local time at 0◦ longitude),
opportunely separating the contribution of dayside and nightside observations. We achieve this by 1. considering
observations in iterative time windows (TW) centered at four MUTs rather than simply at MUT = 12:00 (this is
equivalent to a 6-hour rather than 24-hour moving average), and 2. calculating the local time of each grid point and
imposing that only observations with local times within ±7 hours in the current TW are considered for gridding at that
grid point.
The result of applying this updated methodology is illustrated as an example in Fig. 3 for a sol in the growing phase of
the GDE. In the first column we show the CDOD values effectively used for gridding (specifically for TW = 7 sols) in
the four maps with MUTs = 00:00, 06:00, 12:00, and 18:00. The distinction between nightside tracks (positive slope)
and dayside ones (negative slope) is evident. Because we have used a ±7 hour range for accepting observations at each
grid point, there is a superposition of nightside and dayside values at some longitudes, which allows for a smoother
transition between the two. The distinction in local time is highlighted in the second column of Fig. 3. Here we plot the
local time difference between each observation and the grid point close to which it is located. Clearly there are two
longitude ranges (with local times around 03:00 and 15:00) within which these differences are small, for each different
MUT map. These correspond to the locations spanned by the MRO orbit for that specific MUT, whereas higher values
of the local time difference correspond to locations spanned by the MRO orbit at a different MUT time, but always
within the considered iterative TW. Since we show examples with TW = 7 in Fig. 3, there are multiple orbit tracks with
similar local time differences, but corresponding to different sols. It is the time weight defined in the IWB procedure,
together with the distance and quality of observation weights, that define the contribution of each single observation to
the grid point average, plotted in the fourth column. Note that these gridded maps are already the result of the iterative
application of four time windows -- refer to Figure 6 of Montabone et al. (2015) for an example of how we derive a
gridded map using observations in successive TWs.
It is necessary to discuss in depth the differences among the maps at different MUTs, because these are the novel result
of this work. When looking at the four MUT maps in the fourth row of Fig. 3, in fact, a clear daily variation of CDOD
can be appreciated, particularly pronounced in the latitude band [20◦, 70◦]S (see also Section 3 and Fig. 12). This
variation of CDOD has the characteristics of a Sun-synchronous wave with wavenumber one: smaller optical depths are
found at night, larger optical depths occur during the day. The daily variation is well accounted for in the MCS CDODs,
as shown in the first row of Fig. 3, and is not an artefact of the gridding methodology, nor is limited to the sol showed in
Fig. 3, as Fig. 12 clearly demonstrates. Furthermore, this strong daily variation of CDOD corresponds very well both in
solar longitude (in the growing phase of the GDE) and in latitude to the strong daily variations of the MCS dust opacity
profiles, as described in Kleinböhl et al. [this issue. In that paper, GCM simulations can account for the daily variability
of the dust profiles, and explain the likely dynamical effects at the origin of this phenomenon.
The question arises, then, whether the daily variability observed in MCS CDOD can also have a dynamical origin,
or can be explained otherwise. We address the possibility of a dynamical origin with GCM simulations in Section 4,
while we point out here that interpreting results from MCS CDODs is particularly challenging, as already mentioned
in Subsection 2.2. The third row of Fig. 3, in fact, shows that the dust profiles in the latitude band where the daily
CDOD differences are more pronounced have quite different cut-off altitudes above the local surface between day
and night: nightside profiles tend to extend lower in altitude, while dayside profiles are generally cut-off at higher
altitudes. This is due to several factors, although it is primarily driven by the altitude at which the retrieval algorithm
finds the atmosphere too opaque in the limb path. The increase in the amount of dust (and its vertical extent) in the
dayside profiles causes them to terminate further from the surface than the nightside ones, on average. We have taken
into account the quality of the retrieval fit to the measured radiances by selecting profiles with no more than one
detector rejected due to fitting poorly. As previously pointed out, the different cut-off altitudes for nightside and dayside
retrievals imply that the uncertainty in the CDOD extrapolation is larger during the day, but it does not necessarily
imply that the homogeneously mixed dust assumption is not valid, particularly during the peak of the GDE. We refer to
Section 3 for further considerations about this point.
2.5 Reference MY 34 dust climatology
The gridded and corresponding kriged maps of CDOD described in (Montabone et al., 2015) have been used as reference
multi-annual dust climatology in several studies and applications, including the production of MCD statistics. It is,
therefore, compelling to produce a reference MY 34 climatology following the approach established for the previous
Martian years.
7
A PREPRINT - JULY 19, 2019
Figure 3: This figure shows longitude-latitude maps at four different Mars Universal Times (MUT = 00:00, 06:00,
12:00, and 18:00) in each column, for sol-of-year (SOY) 400, Ls ∼ 196◦, in the growing phase of the GDE (The SOY is
the integer sol number starting from SOY=1 as first sol of the year). In the first row we plot the values of IR absorption
(9.3 µm) CDOD at 610 Pa for observations within a time window TW = 7 sols centered around the corresponding MUT.
In the second row we show the local time difference between the same observations and the grid point around which
they are located. The cut-off altitude above the local surface of the dust profile corresponding to each observation is
shown in the third column. Finally, the gridded maps of IR absorption (9.3 µm) CDOD at 610 Pa obtained by applying
the iterative weighted binning methodology (Montabone et al., 2015) to the observations shown in the first three rows
are plotted in the fourth row. These gridded maps are already the result of the iterative application of four TW = 1, 3, 5,
and 7 sols.
8
A PREPRINT - JULY 19, 2019
Although in this work we produce four gridded maps per sol, we calculate the daily average and we use only one map
per sol to build the reference MY 34 climatology. We do so because 1. the daily variability of MCS CDOD is not
yet soundly confirmed by independent observations, 2. it is not clear that using a dust scenario with daily variability
in model simulations does not trigger spurious effects, e.g. erroneously forcing the tides, and 3. we would like to be
consistent with climatologies from previous Martian years.
We show in Fig. 4 an example of daily averaged gridded map and corresponding kriged one, for the same sol as in
Fig. 3. The gridded map results more complete than any single MUT map, and rather spatially smooth. The transition
to maps at previous and subsequent sols is also rather smooth (see e.g. Figs. 9 and 11). We should mention that, in
contrast to Montabone et al. (2015), we no longer modify the values of the gridded maps in a latitude band around the
southern polar cap edge before applying the kriging interpolation. This was previously done to artificially introduce
climatological "south cap edge storms" and balance TES and MCS years in term of dust lifted at the south cap edge.
The use of MCS v5.3.2 retrievals extending to lower altitudes, and the fact that TES CDOD retrievals at the south cap
edge are being revised (M. Smith, personal communication) alleviates the need for such correction.
The MY 34 daily maps of gridded and kriged IR absorption CDOD normalized to 610 Pa are included in NetCDF files
together with maps of several other variables, as detailed in Appendix B of (Montabone et al., 2015). We note that,
following the Montabone et al. (2015) sol-based Martian calendar (see their Appendix A for a description), MY 34 has
668 sols, therefore we provide 668 gridded maps -- MY 34 new year's solar longitude is 359.98. The dust scenario,
though, has always 669 kriged maps for practical reasons, hence the last sol of the MY 34 dust scenario is the first sol
of MY 35. Both gridded and kriged maps version 2.5 for MY 34 will be publicly available at the dedicated webpage on
the MCD project website hosted by the LMD at the URL: http://www-mars.lmd.jussieu.fr/ after the version of
this manuscript submitted to the Journal of Geophysical Research - Planets is accepted for publication. A beta version
of these maps is currently available upon request to the corresponding author.
2.6 Validation
An important aspect of producing a reference dataset for the dust climatology is its validation with independent
observations. The Opportunity rover entered safe mode right at the onset of the GDE, while the Curiosity rover took
measurements of visible dust optical depth throughout the GDE using its MastCAM camera (Guzewich et al., 2018).
Hence, we use measurements from Curiosity for validation, together with publicly available visible images taken by the
Mars Color Imager (MARCI) camera aboard MRO.
Figure 5 show the comparison between the time series of the dust optical depth (sol-averaged and normalized to 610
Pa) observed by Curiosity rover in Gale crater during the GDE (Guzewich et al., 2018), and the time series of CDOD
extracted from the gridded maps and averaged in a longitude-latitude box centered on Gale crater (after conversion to
equivalent visible values). The gridded maps are able to fairly well reproduce the timing and decay of the GDE around
Gale, but they underestimate the peak of the event. Furthermore, they overestimate the decay between Ls ∼ 205◦ and
Ls ∼ 215◦, although within the uncertainty limit. There are spatial inhomogeneities in the CDOD field even during the
mature phase of the GDE that may account for some of the differences. Also note that Gale crater is a challenging
location for MCS to observe due to MRO providing relay services to the Curiosity rover. In particular the number of
in-track profiles is limited and may be geographically biased. Finally, if kriged maps were used instead of the gridded
maps, the time series would be practically indistinguishable from the magenta line in Fig. 5 (not shown here).
We show the comparison between one of our gridded CDOD maps and a MARCI image in Fig. 6. The comparison is
done for June 6, 2018, at the onset of the GDE, which corresponds to SOY 387 in our dataset. The extension of the dust
cloud in both the MARCI image and the CDOD map is similar, with both showing intense activity around Meridiani, an
eastward progression of the storm, and relatively clear skies over the Tharsis volcanoes. This specific CDOD map fails
to show the onset of the south polar cap edge dust activity, but maps at subsequent sols do.
3 Seasonal, day-to-day, and daily variability of column dust
In this Section we analyze the variability at different temporal scales, which is included in the MY 34 dust climatology
reconstructed from MCS CDODs. In particular, we look at the seasonal, day-to-day, and daily variability, as showed in
Figures 7 to 13.
Starting from the seasonal variability, Fig. 7 shows the latitude vs time plot of the zonally and daily averaged CDOD
obtained from both the gridded maps and the kriged ones. This comparison shows that the kriged maps have the
advantage of being complete (i.e. CDOD values are assigned at every grid point) while preserving the overall properties
of the dust distribution. Montabone and Forget (2018) noted that Martian years show two distinctive seasons with
respect to the atmospheric dust loading, when a comparison of multi-annual zonal means of CDOD is carried out: a
9
A PREPRINT - JULY 19, 2019
Figure 4: Daily averaged gridded map (upper panel) and corresponding kriged map (lower panel) of 9.3 µm absorption
column dust optical depth for SOY 400, Ls ∼ 196◦, in the growing phase of the GDE. The gridded map showed here
is the daily average of the four maps in the fourth column of Fig. 3. The spatial resolution of the gridded map is 6◦
longitude × 5◦ latitude, whereas the resolution of the kriged one is 3◦ longitude × 3◦ latitude. The black boxes in the
maps highlight the averaging area around Gale crater used in Figures 5 for comparison with the CDOD measured by
the Curiosity rover.
10
A PREPRINT - JULY 19, 2019
Figure 5: Time series of equivalent visible column dust optical depth calculated from the 9.3 µm absorption CDOD
normalized to 610 Pa, extracted from the daily averaged gridded maps in an area around Gale crater (magenta line),
compared to the time series of visible column optical depth measured by MastCAM aboard NASA's "Curiosity" rover
(black line). Curiosity observations (Guzewich et al., 2018) have been daily averaged and normalized to 610 Pa (using
the surface pressure from the Mars Climate Database pres0 routine). Both time series are shown between Sol-of-Year
355 and 500, i.e. Ls ∼ [170◦, 260◦]. We used a factor of 2.6 to convert 9.3 µm absorption CDOD into equivalent
visible ones. Data from gridded maps are averaged in the area shown by a black box in Fig. 4 (i.e. longitude=[123◦E,
153◦E], latitude=[15◦S, 10◦N]) centered around Curiosity landing site at longitude 137.4◦E and latitude 4.6◦S. Light
and dark grey shades show the uncertainty envelope (1-sigma) respectively for Curiosity's time series and the time
series extracted from the gridded maps.
"low dust loading" (LDL) season between Ls ∼ 10◦ and Ls ∼ 140◦, and a "high dust loading" (HDL) season at other
times, when regional dust storms and global dust events are most likely to occur -- commonly referred to as the "dust
storm season". MY 34 does not differ, as dust started to increase above the 0.15 level (IR absorption at 9.3 µm) after
Ls ∼ 160◦, following a quiet LDL season (see Fig. 8 as well, which is the time series obtained from the latitude vs
time plot by averaging also in the latitude band [60◦S, 40◦N]).
Nevertheless, the optical depth abruptly increased after Ls ∼ 186◦ due to the onset of the GDE, which rapidly grew to
the west of Meridiani Planum, expanded eastwards and southwards, and spread a large amount of dust at all longitudes
within approximately a latitude band [60◦S, 40◦N] (see its day-to-day evolution over 12 sols in Fig. 9), then slowly
decaying over about 130 sols, as can be observed from the tail of the GDE peak in Fig. 8.
MY 34 also featured two other maxima in CDOD that are climatologically consistent with all other 10 previously
observed years: one at southern polar latitudes centered at Ls ∼ 270◦, and the other one in a latitude band [60◦S, 40◦N]
peaking at Ls ∼ 325◦. These maxima are linked to a regional dust storm occurring over the ice-freed southern polar
region, and to a particularly intense late-winter regional storm (see its day-to-day evolution over 12 sols in Fig.11). The
latter has the characteristics of a flushing storm following the Acidalia-Chryse storm track, although its precise origin
cannot be easily tracked in the gridded maps of CDOD. Finally, the absence of significant storms in a range of solar
longitude centered around Ls ∼ 300◦ -- the so-called "solsticial pause" -- is also climatologically consistent with
what observed in previous years (Montabone et al., 2015; Kass et al., 2016; Lewis et al., 2016; Montabone and Forget,
2018, and Xiaohua et al., this issue).
Before moving to the analysis of the CDOD daily variability, there is a final consideration we ought to do about the
variability of dust storms. When comparing the day-to-day evolution of the GDE and the late winter storm in Figs. 9 and
11 at their early stage, they look pretty similar both in intensity and extension. Furthermore, the shapes of the CDOD
peaks in the time series of Fig. 8 are also comparable (both positively skewed, with sharp increase and long decreasing
tail). What really does make the difference is the fact that a GDE such as the one in MY 34 took about 35 sols of
11
A PREPRINT - JULY 19, 2019
Figure 6: The background global image of Mars in this Figure is referenced PIA22329 in the NASA photojournal
(credits: NASA/JPL-Caltech/MSSS). It shows the growing MY 34 GDE as of June 6, 2018. The map was produced by
the Mars Color Imager (MARCI) camera on NASA's Mars Reconnaissance Orbiter spacecraft. The blue dot shows the
approximate location of the Opportunity rover. We overlap on this image the column dust optical depth kriged map for
the corresponding sol (sol-of-year 387), which we have reconstructed from MCS observations. The IR absorption (9.3
µm) CDOD map (not normalized to 610 Pa) is plotted as filled colored contours.
continuous dust injection into the atmosphere to reach a peak in average CDOD that is more than twice as high than the
one reached by the (still rather intense) late-winter regional storm. This includes a much larger spatial variability during
the GDE, as indicated by the RMSD in Fig. 8. An event that was very important in boosting the equinoctial dust storm
into the GDE class was the activation of secondary lifting centers in the Tharsis region, which seems to start around
SOY 401 in the gridded maps (Ls ∼ 197◦), and later in the Terra Sabaea region -- although one cannot distinguish
from the maps whether the increase of optical depth in this region is the result of eastward transport from Tharsis or
local dust lifting, or both. When looking at the CDOD day-by-day evolution in Fig. 10, this event can be considered as
a second dust storm within the first one, without which we could have just witnessed a regional storm instead of a GDE.
This is one of the reasons why names like "global dust storm" or "planet-encircling dust storm" do not seem to capture
the real nature of this type of extreme events, which are not single storms nor uniquely planet-encircling. Perhaps even
"global dust event" is not particularly appropriate, as high latitude regions are mostly free of dust -- although indirectly
affected by the dust via dynamical effects. One possibility is to call these events for what they really are: "extreme dust
events" (EDE).
Another extreme characteristic of the equinoctial event is its strong daily variability, clearly observed by MCS in the
elevation of the dayside vs nightside dust profiles (see Kleinböhl et al, this issue), but also featured in the column
optical depth values, as already shown in Fig. 3. The time series at different locations extracted from the dataset with
four MUT maps per sol and shown in Fig. 12 clearly illustrates this phenomenon. The nightside-dayside variability is
different at different locations, but is particularly dramatic in Aonia Terra, which is located in the southern latitude band
where Kleinböhl et al [this issue] observe strong variability in the dust profiles. In Fig. 13, therefore, we compare the
CDOD values in Aonia Terra at two times during the GDE (i.e. during its growth phase and near the peak) with the
corresponding dust opacity profiles that are extrapolated and integrated in order to estimate the CDODs. The raise in
altitude of about 20 km (either above local surface or above the areoid) of the dayside dust profiles with respect to the
12
A PREPRINT - JULY 19, 2019
night profiles is quite spectacular at Ls ∼ 207◦, near the peak of the GDE. Unfortunately, with the raise in altitude
of the large opacity values comes the raise in cut-off altitude of the dust profiles. However, from Fig. 13 one cannot
conclude that the homogeneously mixed dust hypothesis at the core of our dust profile extrapolation to the ground does
not hold in these cases. Conversely, there is no evidence that raising dust is replaced by more well mixed dust in the
missing part of the profile, because we simply have no data. Furthermore, uncertainties at the lowest levels of dust
profiles during the GDE tend to be larger (see e.g. left panel of Fig. 1 in Kleinböhl et al. this issue), hence the real
shape of the profile in the lowest two scale heights could provide some surprise.
At this point of the analysis, we can make three hypotheses about the daily variability observed in MCS CDOD:
1. There is an intrinsic variability of the column dust. In this case, quite a substantial amount of dust must be
supplied in the lowest two scale heights during the day, where MCS cannot see through. This extra dust must
either be lifted locally from the ground or supplied via horizontal advection (or both). Local mesoscale effects
might operate at different locations (e.g. katabatic/anabatic winds, strong convective activity, etc.)
2. There is no significant variability in the column dust. In this case, atmospheric dust is simply moved up and
down during the day/night, and the dayside dust opacities actually decrease with decreasing altitude in the
lowest scale heights, not seen by MCS. The daily variability of the dust opacity profiles in the lowest two
scale heights during the GDE would then be expected to be very large, in order to compensate for the raise in
altitude of the dust cloud.
3. There is some variability in the column dust. In this case, dust is partly moved up and down at different local
times, and partly lifted locally, or advected from nearby locations.
In order to help us clarifying which hypothesis is more likely, we have carried out simulations with the LMD-GCM,
which we are going to discuss in the next section.
Figure 7: MY 34 latitude vs time plot of the zonally and daily averaged gridded maps of 9.3 µm absorption column
dust optical depth normalized to the reference pressure level of 610 Pa (upper panel), compared to the same using
kriged maps (lower panel). White colour in the upper panel indicates missing data.
13
A PREPRINT - JULY 19, 2019
Figure 8: Time series of column dust optical depth (9.3 µm in absorption, normalized to 610 Pa) extracted from the
daily averaged gridded maps and averaged at all longitude in the latitude band [60◦S, 40◦N]. The grey shade represents
the root mean squared deviation, i.e. the spatial variability within the averaged longitudes and latitudes (note that the
daily variability is not included).
4 Global Climate Model simulations of the MY 34 Global Dust Event
The simulations we have carried out using the LMD-MGCM have similar characteristics to those carried out to build
the Mars Climate Database version 5.3 (Millour et al., 2015), except for the model top being set at 100 km (with 29
vertical levels) and the thermospheric parameterizations (González-Galindo et al., 2011) being switched off. The most
up-to-date physical parameterizations are included: interactive dust cycle (Madeleine et al., 2011), thermal plume model
(Colaïtis et al., 2013), water cycle with radiative effect of clouds (Madeleine et al., 2012) and full microphysics scheme
(Navarro et al., 2014). The "rocket dust storm" parameterization recently built and tested by Wang et al. (2018) is not
included in this version of the GCM. The horizontal grid features 64 × 48 longitude-latitude points.
The initial state for the MY 34 run at Ls = 0◦ uses the "climatological" dust scenario typical of MYs devoid of global
dust events. Then, two types of simulations are carried out:
1. a simulation using the reference MY 34 dust scenario v2.5 (i.e. the maps kriged from the daily averaged
gridded maps, as discussed in Section 2.5) to force the dust column field throughout the GDE period;
2. a simulation using the MY 34 dust scenario until Ls = 210◦ (around the peak of the GDE), then continuing as
a "free-dust" run for a few sols, with no more external forcing on the column dust field.
Furthermore, given the availability of a dust scenario for MY 25 (Montabone et al., 2015) and the presence of an
equinoctial GDE in that year, with comparable characteristics to the one in MY 34, we have also carried out a forced
simulation for MY 25, for comparison.
The LMD-MGCM simulations for MY 34 are also used in Kleinböhl et al. [this issue] to discuss the diurnal cycle of the
vertical distribution of dust observed by MCS. We use the two types of simulations for two different purposes: 1. the
forced run is used to analyse some of the impacts of the MY 34 GDE on the local and global dynamics, thus verifying
that the use of a daily averaged dust scenario produces reasonable results, consistent with those for MY 25; 2. the "free
dust" run is used to identify possible daily variability of the column dust in the model, which could corroborate one of
the three hypotheses provided at the end of last Section,
Figure 14 shows a comparison between the surface temperature measured by the Mars Science Laboratory (MSL
Curiosity Guzewich et al., 2018) and the surface temperature computed by the LMD Mars GCM. When the MY 34
global dust event starts, the diurnal amplitude of temperature is reduced: daytime temperatures are lower as a result of
14
A PREPRINT - JULY 19, 2019
Figure 9: Initial evolution of the MY 34 Global Dust Event. Each panel shows daily averaged gridded column dust
optical depth (in absorption at 9.3µm) normalized to the reference pressure level of 610 Pa. From top left to bottom
right, maps are provided from sol-of-year 383 to sol-of-year 394 (i.e. from Ls ∼ 186◦ to Ls ∼ 193◦). See also
Appendix A of Montabone et al. (2015) for the description of the sol-based Martian calendar we use in this paper.
15
A PREPRINT - JULY 19, 2019
Figure 10: Same as Fig.9 but for the MY 34 secondary storm within the GDE. From top left to bottom right, maps
are provided from sol-of-year 401 to sol-of-year 412 (i.e. from Ls ∼ 197◦ to Ls ∼ 204◦). Note that the scale for the
CDOD values has changed with respect to Fig. 9.
16
A PREPRINT - JULY 19, 2019
Figure 11: Same as Fig.9 but for the MY 34 late-winter regional storm. From top left to bottom right, maps are provided
from sol-of-year 596 to sol-of-year 607 (i.e. from Ls ∼ 320◦ to Ls ∼ 326◦). The scale for the CDOD value is the same
as in Fig. 9.
17
A PREPRINT - JULY 19, 2019
Figure 12: Time series of column dust optical depth (9.3 µm in absorption, normalized to 610 Pa) extracted from
the gridded maps with four MUT per sol in three different areas: Meridiani Planum (longitude=[15◦W, 15◦E],
latitude=[15◦S, 15◦N), Hellas Planitia (longitude=[55◦E, 85◦E], latitude=[60◦S, 30◦S), and Aonia Terra (East of
Argyre Planitia: longitude=[90◦W, 60◦W], latitude=[60◦S, 30◦S). The time series are shown between Sol-of-Year 355
and 500, i.e. Ls ∼ [170◦, 260◦].
visible absorption of incoming sunlight being more efficient in a dustier atmosphere, and nighttime temperatures are
higher as a result of increased infrared radiation emitted towards the surface in a dustier atmosphere. The temporal
variability of temperature (absolute and relative values) is well reproduced for the nighttime minimum temperature, but
less so for the daytime temperatures (although the qualitative behaviour is correct). There might be three reasons for
this: 1. thermal inertia is not well represented in the LMD-MGCM for daytime conditions in Gale Crater; 2. the CDOD
observed by MCS in the region of Gale crater is underestimated with respect to the one observed by Curiosity from
Ls ∼ 195◦ to Ls ∼ 202◦, and by consequence the corresponding gridded maps and dust scenario are low-biased at
those times (see Fig. 5 in Section 2.6); 3. the accuracy of the calculations by the model radiative transfer could decrease
under extreme dust loading conditions, or could be affected by an inaccurate distribution of particle sizes.
An important test of the dynamical behavior of our LMD-MGCM simulation forced by the MY 34 dust scenario is how
thermal tides react to the global increase of dust opacity following the onset of the GDE. Figure 15 shows several daily
cycles of surface pressure corresponding to the onset of the MY 25 and MY 34 GDEs; Figure 16 presents a spectral
analysis. Both the amplitude of the daily pressure cycle, and its morphology, are modified by the GDE at its onset. The
diurnal pressure cycle is dominated by the diurnal tide before the GDE takes place. When the GDE starts to build up
and the column optical depth increases, the diurnal mode increases slightly in amplitude while the semi-diurnal mode
increases significantly compared to the other modes, as is already described in previous studies (Zurek and Martin,
1993; Wilson and Hamilton, 1996; Lewis and Barker, 2005, their Figure 5). The reinforcement of the semi-diurnal tide
with increased column opacity is due to the fact that this tide component is dominated by a Hough mode with a large
vertical wavelength (Chapman and Lindzen, 1970). As a result, this Hough mode is very sensitive to forcing extended
in altitude such as the absorption of incoming sunlight by dust particles during a dust storm.
Those major changes in the tidal modes take only a couple of sols to react to the GDE onset: after their respective
onsets, the GDE of MY 34 and MY 25 basically exhibit a very similar diurnal pressure cycle. This similarity between
MY 34 and MY 25 is actually a good cross-calibration test for the reconstructed dust scenarios, since MY 25 is based
on retrieved column optical depths from TES nadir observations, while MY 34 is based on estimated values from MCS
limb observations. Interestingly, comparing the cases of MY 25 and MY 34 at the Curiosity and Opportunity landing
sites in Figure 15 highlights the fact that the modification of the thermal tides is only occurring locally at the site at
which the column optical depth started to rise, and not at the other site. This indicates that the extent of the storm has to
become large enough for the global modification of the thermal tide to be effective.
18
A PREPRINT - JULY 19, 2019
Figure 13: Data plotted in each panel of this figure are for 7 sols centered on either sol-of-year 400 (Ls ∼ [196◦,
upper panels) or sol-of-year 418 (Ls ∼ [207◦, lower panels) in Aonia Terra (longitude=[90◦W, 60◦W], latitude=[60◦S,
30◦S). Blue indicate nightside data, red is for dayside data. The left panels of this figure shows the MCS CDOD in
extinction at 21.6 µm (not normalized to 610 Pa) as a function of the cut-off altitudes of their corresponding dust
opacity profiles, which are plotted in the central panels as a function of altitude above local surface, and in the right
panels as a function of altitude above the areoid (topography values are interpolated from the MOLA dataset at the
corresponding longitudes and latitudes). Note that the x-axis and y-axis ranges are different among the panels.
The increase in column dust optical depth associated with the MY 34 GDE has a profound impact on the large-scale
circulation. Lewis and Read (2003) evidenced an equatorial low-troposphere super-rotating jet in the atmosphere of
Mars, and emphasized the strong positive impact of the atmospheric dust loading on this jet. Our LMD-MGCM forced
simulation for MY 34 shows that the intensity of this super-rotating jet (diagnosed by the super-rotating index as in
Lewis and Read, 2003) is indeed increased following the onset of the GDE from a 5% super-rotation index to a 15%
super-rotation index. We also find that this jet is becoming more confined close to the surface as the GDE develops
(Figure 17 top). The mean meridional circulation is also deeply impacted by the large dust loading following the onset
19
A PREPRINT - JULY 19, 2019
Figure 14: Comparison of surface temperature simulated by the LMD-MGCM model versus surface temperature
measured by the Rover Environmental Monitoring Station on board MSL "Curiosity" rover (daily minimum in blue and
daily maximum in red). Data from MSL are provided as supplementary material of Guzewich et al. (2018)
Figure 15: Daily cycle of surface pressure as simulated by the LMD Mars GCM using the MY 25 and MY 34 dust
scenarios. The focus of the figure corresponds to the onset of the two GDEs. This is showing the simulated fields at the
Opportunity (left) and Curiosity (right) landing sites.
of the MY 34 GDE: the intensity of this mean meridional circulation is enhanced by a factor of 10 following the onset
and mature phase of the GDE (Figure 17 bottom). This behaviour is similar to the evolution of the mean meridional
circulation simulated under MY 25 GDE conditions (e.g., Montabone et al., 2005).
Finally, we discuss the use of a "free dust" simulation to get some insights on the daily variability of CDOD in the
GCM model. GCM simulations have already proved successful to show that large-scale circulation components (i.e. the
mean meridional circulation, planetary waves, and the polar vortex) cause the vertical distribution of dust to undergo
diurnal variations both in equatorial and extratropical regions [Kleinböhl et al., this issue]. Figure 18 shows the column
dust optical depth simulated in a LMD-MGCM run around Ls = 210◦ (i.e. near the peak of the MY 34 storm).
The total column optical depth freely evolves in the simulation without being re-normalized using the values in the
MY 34 dust scenario. As is observed in the MCS CDOD values, and by consequence in the gridded/kriged maps
reconstructed following the method described in this paper, the column optical depth in the "free dust" model run
could vary significantly on a diurnal basis in a given region. This results from horizontal transport by the large-scale
circulation. While it is not possible to rule out the other possible interpretations of the observed diurnal variability of
column optical depth as discussed at the end of Section 3, the LMD-MGCM results in Figure 18 strongly suggests
that this variability has a physical basis, and at least part of it would be related to the large-scale horizontal transport.
What our GCM simulation cannot tell is how much specific mesoscale phenomena, including dusty deep convection
("rocket dust storms" Spiga et al., 2013), or Planetary Boundary Layer processes, would contribute to the stronger
20
170180190200210220230240250Ls (degrees)165180195210225240255270285300Surface temperature (K)LMD GCMMSL minMSL max180182184186188190192194196198200Ls0.00.81.62.43.24.04.85.66.47.28.0 at 610 PaMY25MY34180182184186188190192194196198200Ls0.00.81.62.43.24.04.85.66.47.28.0 at 610 PaMY25MY34180182184186188190192194196198200Ls510525540555570585600615630645Surface pressure (Pa)MY25MY34180182184186188190192194196198200Ls540555570585600615630645660675Surface pressure (Pa)MY25MY34A PREPRINT - JULY 19, 2019
Figure 16: Planetary waves detected by a Fast Fourier transform performed on the surface pressure simulated at the
equator by the LMD Martian GCM forced by the MY 34 dust scenario. The spectral analysis is carried out on two time
intervals before the MY34 GDE (sol-of-year 350 to 370, left) and during the MY 34 GDE (sol-of-year 400 to 420, right).
The diurnal (-1), semi-diurnal (-2) and quarter-diurnal (-4) tides are detected, as well as the Kelvin (+1) mode. The
most notable change is the reinforcement of the semi-diurnal mode by the MY 34 GDE, with an enhancement of spectral
power of about one order of magnitude. During the GDE, the semi-diurnal tidal mode is as strong as the diurnal tidal
mode. The quarter-diurnal mode is also significantly reinforced, becoming stronger than the wavenumber-1 Kelvin
mode.
diurnal variability observed by MCS. It should also be pointed out that the GCM model uses a simplified scheme for
dust lifting, which might not be well suited to reproduce strong day-night variability.
5 Conclusions and remarks
The work described in this paper was devoted 1. to reconstruct maps of column dust optical depth for MY 34 from
Mars Reconnaissance Orbiter/Mars Climate Sounder observations, 2. to analyze the seasonal, day-to-day, and daily
variability of column dust showed by the maps, and 3. to use numerical simulations with the Laboratoire de Météorologie
Dynamique Mars global climate model, forced by or simply initiated with the reconstructed CDOD maps, in order to
examine some aspects of the impact of the MY 34 global dust event on local and global scale dynamics, including the
daily variability of column dust.
The reconstructed maps for MY 34 follow the work by Montabone et al. (2015) and extend the publicly available
multi-annual, multi-instrument climatology of column dust optical depth to 11 Martian years. An important difference
of the present work with respect to Montabone et al. (2015) is that we have now reconstructed sub-daily maps of column
dust, allowing us to access the analysis of the daily variability of such quantity. This was made possible by using novel
retrievals (version 5.3.2) of dust opacity profiles from MCS observations during the period May 21, 2018, to October
15, 2018 (Ls = [179◦, 269◦] in MY 34), which extend lower in altitude than standard version 5.2 retrievals. In general,
therefore, the estimated column dust optical depth values during the global dust event of MY 34 result more accurate,
within the intrinsic limitations of estimating CDODs from limb observations.
The analysis of the MY 34 column dust variability at different temporal scales using the reconstructed maps highlighted
that:
• MY 34 reproduced the dichotomy observed in the 10 previous years between the "low dust loading" and the
"high dust loading" seasons (this might not be the case for the recently started MY 35, which featured an
unusually intense dust storm during the LDL season, D. Kass, personal communication);
• It also featured other typical characteristics of the seasonal evolution of CDOD, such as large values at southern
polar latitudes peaking at Ls ∼ 270◦, a solsticial pause centered at Ls ∼ 300◦, and large values peaking again
at Ls ∼ 325◦ during the evolution of an intense late-winter dust storm;
21
A PREPRINT - JULY 19, 2019
Figure 17: The impact of the MY 34 GDE on the zonally-averaged global circulations on Mars is shown from
left to right, averaged on the Ls intervals [150◦, 180◦] (pre-GDE conditions), [180◦, 210◦] (onset of the GDE), and
[210◦, 240◦] (mature phase of the GDE). [Top] Super-rotation index s computed according to Lewis and Read (2003)
with positive values denoting regions where eastward jets are super-rotating i.e. exceeding the solid-body rotation of
the planet. [Bottom] Mass streamfunction with blue regions corresponding to counterclockwise circulation and red
regions corresponding to clockwise circulation.
• The key distinction of MY 34 was undoubtedly the equinoctial global dust event (starting at Ls ∼ 186◦
only few sols after the equivalent event in MY 25), which seemed to feature a "storm within the storm" at
Ls ∼ 197◦, boosting its growth to attain extreme characteristics, typical of GDEs;
• The MY 34 GDE seems also to feature very large CDOD diurnal variability at selected locations, particularly
at southern mid- and high-latitudes, as already observed in the corresponding MCS dust opacity profiles by
Kleinböhl et al. [this issue].
22
A PREPRINT - JULY 19, 2019
Figure 18: Sequence of column dust optical depth maps separated by 3 hours over two sols in a LMD-MGCM "free
dust" simulation where, contrary to the simulation forced with the MY 34 dust scenario, the dust mass mixing ratio in
the model is not normalized to match the total column dust optical depth in the MY 34 dust scenario. This "free dust"
simulation was restarted from the simulated state of the atmosphere at Ls = 210◦ in a regular LMD-MGCM simulation
forced by the MY 34 dust scenario.
23
A PREPRINT - JULY 19, 2019
While the observation of the diurnal variability in dust opacity profiles comes from direct MCS retrievals, and could be
explained by global climate model simulations invoking the effects of the large-scale circulation [Kleinböhl et al., this
issue], the same observation in the indirectly estimated column dust values poses more challenging questions. Is the
daily variability intrinsic to the column dust, or should we expect that the shape of the dust profile in the lowest one
or two scale heights not directly observed by MCS (particularly in dayside observations) is not compatible with an
homogeneously mixed assumption? Whether the answer leans towards the former or the latter, or a bit of both, it would
open a novel view of how dust is three-dimensionally distributed within Martian dust storms.
It is not the purpose of this paper to provide a definite answer to the aforementioned question. Nevertheless, we have
resorted to numerical simulations with the LMD-MGCM to provide us with some hints. Using a "free dust" model run
initiated at Ls=210◦ with the reconstructed CDOD field, the model is able to reproduce some daily variability in column
dust at selected locations. Despite the fact that both the range of the variability and the precise locations do not coincide
with what is estimated from MCS, this result provides a physical evidence that some degree of daily variability can be
expected not only in the dust profiles but also in the integrated columns. Furthermore, the dust lifting and planetary
boundary layer parameterizations of the global model might actually miss some of the important features that lead to an
accurate description of the three-dimensional dust distribution. The model result might, therefore, underestimate the
real variability of the column dust.
What the model simulations clearly show when forced with daily averaged CDOD maps, however, is that the impact of
the MY 34 GDE on the atmospheric dynamics is as large as for the MY 25 GDE. Key features of the local and global
dynamics (such as tides, mean meridional circulation, and equatorial winds) respond to the equinoctial dust events
in a very similar manner. This is also an indirect validation of the MY 34 reference dust scenario based on the daily
averaged CDOD maps from MCS, which is currently used in several modeling studies of the 2018 GDE.
Future work should address the possibility of producing and making available a sub-daily dust scenario. As mentioned
in Subsection 2.5, we consider that this option was not currently viable, mainly because the CDOD daily variability is
not yet independently confirmed, and because it is not yet clear whether model simulations forced by a sub-daily dust
scenario are free of spurious effects. Current Mars GCMs might need to be adapted to handle diurnally varying CDODs
in a stable and sensible fashion, if the degree of variability will be proved to be of the order of the one showed in this
paper.
Strong emphasis should also be put in obtaining future observations of column-integrated dust as well as dust profiles
with sub-daily frequency. The Planetary Fourier Spectrometer (PFS) aboard Mars Express, and the Atmospheric
Chemistry Suite (ACS) aboard Trace Gas Orbiter currently provide the capability to retrieve CDOD at multiple
local times, and could help in the comparison with the estimated MCS day-night variability of column dust. Future
observations from the forthcoming Emirates Mars Mission (EMM) might provide even stronger evidence of the presence
or absence of daily variability.
Moreover, in order to fully characterize the diurnal cycle of dust and accurately monitor the evolution of dust storms on
Mars, novel approaches must be taken in the future. These include:
• the use of satellites in Mars-stationary orbits (also called "areostationary"), which are equatorial, circular,
planet-synchronous orbits equivalent to geostationary ones for the Earth (see e.g. Montabone et al., 2018, ;)
• the use of instruments that allow to observe the vertical distribution of the dust in the Martian PBL, such as
lidars. This is particularly important during dust storms when IR spectrometers/radiometers (both nadir- and
limb-looking) fail to produce reliable retrievals because of the large atmospheric opacity and the reduced
temperature contrast.
A Appendix: MY 34 climatology versioning
MCS version 5.3.2 is an interim, experimental version of retrievals, leading to a possible future improved version
of the whole MCS dataset. As a consequence, the MY 34 gridded and kriged datasets should be considered work
in progress, as should the datasets related to other Martian years. It is our intention to regularly update the multi-
annual, multi-instrument dust climatology with new observations, novel retrievals of past observations, and updated
gridding methodologies/features. The updates are likely to be made publicly available on the Mars Climate Database
project webpage at the URL http://www-mars.lmd.jussieu.fr/. There exist, therefore, multiple versions of these
reference dust climatologies, notably for MY 34, and we would like to provide in this appendix some details about the
main differences.
As mentioned in Section 2, the reference version of the current maps from MY 24 to 32 is v2.0. For this version,
the used gridding/kriging methodology is precisely the one described in Montabone et al. (2015). Version 2.1 is a
24
A PREPRINT - JULY 19, 2019
specific version only for MY 33, where we have used a different weight for THEMIS observations with respect to MCS
observations, in order to account for THEMIS retrievals provided at later local times, and we started using MCS v5.2
"two-dimensional" retrievals (Kleinböhl et al., 2017a) instead of v4.3 "one-dimensional" retrievals for previous years 28
to 32.
For MY 34, we have produced three intermediate versions (v2.2, v2.3, and v2.4) and the v2.5 described in this paper,
which should be considered as the reference version. All three intermediate versions use MCS v5.3.2 retrievals for
the available period, and MCS v5.2 for the rest of the time, but do not use the two distinctive features described in
Subsection 2.4, namely the local time cut-off window of ±7 hours for observations considered for the weighted average
at each grid point, and the 6 hour moving average producing 4 maps per sol. Instead, they use observations at all local
times for each grid point (except in v2.3 and v2.4 during the GDE, see below), and the 24 h moving average produces
only one map per sol, centered at MUT=12:00, as described in Montabone et al. (2015).
Within the intermediate versions, the differences are as following:
• v2.2: This version still uses the same data QC and gridding methodology as in Montabone et al. (2015). The
use of dayside values is limited by the application of the "dayside" filter with 8 km cut-off altitude threshold at
any time. Apart the use of MCS v5.3.2 retrievals, the only other difference with respect to v2.0 is in the kriged
maps, where we have artificially introduced climatological south cap edge "storms" only for solar longitude
earlier than 180◦, as in MY 25. This version also uses MCS observations only until end of September 2018
(Ls ∼ 260◦), stopping at SOY 501.
• v2.3: This version uses only dayside values during the GDE (186.5◦ < Ls < 269◦, SOY 383 to 515). It has
also an improved data QC with respect to v2.2: we introduced the "water ice" filter, the "cross-track" filter, and
we did not apply the "dayside filter" with 8 km cut-off altitude threshold during the GDE and the late winter
regional storm (Ls > 312◦). This allowed to use many more dayside values during the two major dust events
of MY 34, increasing the overall optical depth to levels observed by, e.g., Opportunity rover. We also redefined
the estimation of uncertainties according to the scheme that was later adopted in v2.5 (but with slightly lower
uncertainties overall). Furthermore, we changed a couple of parameters in the IWB methodology: the criterion
to accept a value of weighted average at a particular grid point at any given iteration became that there must be
at least one observation within a distance of 200 km from the grid point. We started using the same surface
pressure recorded in the MCS dataset to renormalize CDOD to 610 Pa, instead of the MCD surface pressure.
If MCS surface pressure is not retrieved, we associated a 10% uncertainty by default. We stopped using the
artificial modification of a latitude band around the southern polar cap at all times. This version also uses MCS
observations only until end of February 2018 (Ls ∼ 349◦), stopping at SOY 647.
• v2.4: This version is quite similar to v2.3. The only differences are in the refined data QC, which is the one we
also use in v2.5 (see Subsection 2.2). It also extends until the end of MY 34.
Refer to Figures 19 and 20 for a comparison of results using versions 2.2, 2.3, 2.4 and 2.5.
Acknowledgments
The work related to the production of reconstructed gridded and kriged CDOD maps for MY 34 is funded by the French
Centre National d'Etudes Spatiales (CNES). This work uses technical achievements and expertise obtained during a
parallel project of improving the gridding of column dust optical depth retrievals from satellite observations, funded by
NASA PDART programme (Grant no. NNX15AN06G).
The maps of gridded and kriged CDOD produced in this work will be publicly available under the Mars Climate
Database project (at the URL http://www-mars.lmd.jussieu.fr/) after the version of this manuscript submitted
to the Journal of Geophysical Research - Planets is accepted for publication. A beta version of these maps is currently
available upon request to the corresponding author ([email protected]). Mars Climate Sounder data is
publicly available on NASA's Planetary Data System (https://pds-atmospheres.nmsu.edu/). Work related to MCS
observations and retrievals (including the estimates of CDOD) is carried out at the Jet Propulsion Laboratory, California
Institute of Technology, and is performed under a contract with NASA. Government sponsorship is acknowledged. Data
from Curiosity rover obtained by the Rover Environmental Monitoring Station (REMS) instrument is publicly available
as supplementary material of Guzewich et al. (2018).
The authors wish to thank R. John Wilson for useful comments on early versions of the reference MY 34 dust
climatology datasets.
Luca Montabone wishes to dedicate this work to his father Augusto, who started his journey to the stars during the
production of this paper.
25
A PREPRINT - JULY 19, 2019
References
B. A. Cantor. MOC observations of the 2001 Mars planet-encircling dust storm. Icarus, 186:60 -- 96, 2007. doi:
10.1016/j.icarus.2006.08.019.
S. Chapman and R. Lindzen. Atmospheric tides. Thermal and gravitational. Dordrecht: Reidel, 1970, 1970.
A. Colaïtis, A. Spiga, F. Hourdin, C. Rio, F. Forget, and E. Millour. A thermal plume model for the Martian convective
boundary layer. Journal of Geophysical Research (Planets), 118:1468 -- 1487, 2013. doi: 10.1002/jgre.20104.
A. Fedorova, J.-L. Bertaux, D. Betsis, F. Montmessin, O. Korablev, L. Maltagliati, and J. Clarke. Water vapor in the
middle atmosphere of Mars during the 2007 global dust storm. Icarus, 300:440 -- 457, 2018. doi: 10.1016/j.icarus.
2017.09.025.
F. Forget, A. Spiga, B. Dolla, S. Vinatier, R. Melchiorri, P. Drossart, A. Gendrin, J.-P. Bibring, Y. Langevin, and
B. Gondet. Remote sensing of surface pressure on Mars with the Mars Express/OMEGA spectrometer: 1. Retrieval
method. Journal of Geophysical Research (Planets), 112(E11):8 -- +, 2007. doi: 10.1029/2006JE002871.
F. González-Galindo, A. Määttänen, F. Forget, and A. Spiga. The martian mesosphere as revealed by co2 clouds
observations and general circulation modeling. Icarus, 216:10 -- 22, 2011.
Scott D. Guzewich, M. Lemmon, C. L. Smith, G. Martínez, A. de Vicente-Retortillo, C. E. Newman, M. Baker,
C. Campbell, B. Cooper, J. Gómez-Elvira, A.-M. Harri, D. Hassler, F. J. Martin-Torres, T. McConnochie, J. E.
Moores, H. Kahanpää, A. Khayat, M. I. Richardson, M. D. Smith, R. Sullivan, M. de la Torre Juarez, A. R.
Vasavada, D. Viúdez-Moreiras, C. Zeitlin, and Maria-Paz Zorzano Mier. Mars science laboratory observations of the
2018/mars year 34 global dust storm. Geophysical Research Letters, 0(ja), 2018. doi: 10.1029/2018GL080839. URL
https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2018GL080839.
N. G. Heavens, A. Kleinböhl, M. S. Chaffin, J. S. Halekas, D. M. Kass, P. O. Hayne, D. J. McCleese, S. Piqueux,
J. H. Shirley, and J. T. Schofield. Hydrogen escape from Mars enhanced by deep convection in dust storms. Nature
Astronomy, 2:126 -- 132, 2018. doi: 10.1038/s41550-017-0353-4.
Melinda A. Kahre, James R. Murphy, Claire E. Newman, R. John Wilson, Bruce A. Cantor, Mark T. Lemmon, and
Michael J. Wolff. The Mars Dust Cycle, pages 229 -- 294. 2017. doi: 10.1017/9781139060172.010.
D. M. Kass, A. Kleinböhl, D. J. McCleese, J. T. Schofield, and M. D. Smith. Interannual similarity in the Martian
atmosphere during the dust storm season. Geophysical Research Letters, 43(12):6111 -- 6118, Jun 2016. doi:
10.1002/2016GL068978.
A. Kleinböhl, J. T. Schofield, D. M. Kass, W. A. Abdou, C. R. Backus, B. Sen, J. H. Shirley, W. G. Lawson, M. I.
Richardson, F. W. Taylor, N. A. Teanby, and D. J. McCleese. Mars Climate Sounder limb profile retrieval of
atmospheric temperature, pressure, and dust and water ice opacity. Journal of Geophysical Research (Planets), 114:
E10006, October 2009. doi: 10.1029/2009JE003358.
A. Kleinböhl, J. T. Schofield, W. A. Abdou, P. G. J. Irwin, and R. J. de Kok. A single-scattering approximation for
infrared radiative transfer in limb geometry in the Martian atmosphere. J. Quant. Spectrosc. Rad. Transfer, 112:
1568 -- 1580, 2011.
A. Kleinböhl, L. Chen, and J. T. Schofield. Far Infrared Spectroscopic Parameters of Mars Atmospheric Aerosols and
their Application to MCS Retrievals in High Aerosol Conditions. In F. Forget and M. Millour, editors, The Mars
Atmosphere: Modelling and observation, page 2230, January 2017b.
Armin Kleinböhl, R. John Wilson, David Kass, John T. Schofield, and Daniel J. McCleese. The semidiurnal tide in the
middle atmosphere of Mars. Geophysical Research Letters, 40(10):1952 -- 1959, May 2013. doi: 10.1002/grl.50497.
A. Kleinböhl, A. J. Friedson, and J. T. Schofield. Two-dimensional radiative transfer for the retrieval of limb emission
measurements in the martian atmosphere. Journal of Quantitative Spectroscopy and Radiative Transfer, 187:511 --
522, 2017a. ISSN 0022-4073. doi: https://doi.org/10.1016/j.jqsrt.2016.07.009. URL http://www.sciencedirect.
com/science/article/pii/S0022407316302667.
J. S. Levine, D. Winterhalter, and R. L. Kerschmann. Dust in the Atmosphere of Mars and its Impact on Human
Exploration. Cambridge Scholars Publishing, UK, 2018.
S. R. Lewis and P. R. Barker. Atmospheric tides in a Mars general circulation model with data assimilation. Advances
in Space Research, 36:2162 -- 2168, 2005. doi: 10.1016/j.asr.2005.05.122.
S. R. Lewis and P. L. Read. Equatorial jets in the dusty Martian atmosphere. Journal of Geophysical Research (Planets),
108:5034 -- +, 2003. doi: 10.1029/2002JE001933.
S. R. Lewis, D. P. Mulholland, P. L. Read, L. Montabone, R. J. Wilson, and M. D. Smith. The solsticial pause on Mars:
1. A planetary wave reanalysis. Icarus, 264:456 -- 464, 2016. doi: 10.1016/j.icarus.2015.08.039.
26
A PREPRINT - JULY 19, 2019
J.-B. Madeleine, F. Forget, E. Millour, L. Montabone, and M. J. Wolff. Revisiting the radiative impact of dust on Mars
using the LMD Global Climate Model. Journal of Geophysical Research (Planets), 116:E11010, November 2011.
doi: 10.1029/2011JE003855.
J.-B. Madeleine, F. Forget, E. Millour, T. Navarro, and A. Spiga. The influence of radiatively active water ice clouds on
the Martian climate. Geophys. Res. Lett., 39:L23202, 2012. doi: 10.1029/2012GL053564.
M. C. Malin, B. A. Cantor, and Britton A. W. Mro marci weather report for the week of 4 june 2018 -- 10 june 2018.
Malin Space Science Systems Captioned Image Release, MSSS-534, http://www.msss.com/msss_images/
2018/06/13/, 2018.
Leonard J. Martin and Richard W. Zurek. An analysis of the history of dust activity on Mars. J. Geophys. Res., 98(E2):
3221 -- 3246, 1993.
D. J. McCleese, J. T. Schofield, F. W. Taylor, S. B. Calcutt, M. C. Foote, D. M. Kass, C. B. Leovy, D. A. Paige, P. L.
Read, and R. W. Zurek. Mars Climate Sounder: An investigation of thermal and water vapor structure, dust and
condensate distributions in the atmosphere, and energy balance of the polar regions. Journal of Geophysical Research
(Planets), 112(E11):5 -- +, 2007. doi: 10.1029/2006JE002790.
E. Millour, F. Forget, A. Spiga, T. Navarro, J.-B. Madeleine, L. Montabone, A. Pottier, F. Lefevre, F. Montmessin,
J.-Y. Chaufray, M. A. Lopez-Valverde, F. Gonzalez-Galindo, S. R. Lewis, P. L. Read, J.-P. Huot, M.-C. Desjean,
and MCD/GCM development Team. The Mars Climate Database (MCD version 5.2). European Planetary Science
Congress 2015, 10:EPSC2015 -- 438, 2015.
L. Montabone and F. Forget. Forecasting Dust Storms on Mars: a Short Review, page 000. Cambridge Scholars
Publishing, Ed. Levine, G. S. and Winterhalter, D. and Kerschmann, R. L., 2018. doi: doi_tbd.
L. Montabone, S. R. Lewis, and P. L. Read. Interannual variability of Martian dust storms in assimilation of several
years of Mars global surveyor observations. Advances in Space Research, 36:2146 -- 2155, 2005. doi: 10.1016/j.asr.
2005.07.047.
L. Montabone, F. Forget, E. Millour, R. J. Wilson, S. R. Lewis, B. Cantor, D. Kass, A. Kleinböhl, M. T. Lemmon,
M. D. Smith, and M. J. Wolff. Eight-year climatology of dust optical depth on Mars. Icarus, 251:65 -- 95, 2015. doi:
10.1016/j.icarus.2014.12.034.
Luca Montabone, Francois Forget, Michael Smith, Bruce Cantor, Michael Wolff, Michel Capderou, and Michael
VanWoerkom. Mars Aerosol Tracker (MAT): An Areostationary SmallSat to Monitor Dust Storms and Water Ice
Clouds. In 42nd COSPAR Scientific Assembly, volume 42, pages B0.2 -- 16 -- 18, Jul 2018.
D. P. Mulholland, P. L. Read, and S. R. Lewis. Simulating the interannual variability of major dust storms on Mars
using variable lifting thresholds. Icarus, 223:344 -- 358, 2013. doi: 10.1016/j.icarus.2012.12.003.
T. Navarro, J.-B. Madeleine, F. Forget, A. Spiga, E. Millour, F. Montmessin, and A. Määttänen. Global Climate
Modeling of the Martian water cycle with improved microphysics and radiatively active water ice clouds. Journal of
Geophysical Research (Planets), 2014. doi: 10.1002/2013JE004550.
J. A. Ryan and R. M. Henry. Mars atmospheric phenomena during major dust storms as measured at surface.
J. Geophys. Res., 84:2821 -- 2829, 1979.
A. Spiga, J. Faure, J.-B. Madeleine, A. Määttänen, and F. Forget. Rocket dust storms and detached dust layers in the
Martian atmosphere. Journal of Geophysical Research (Planets), 118:746 -- 767, April 2013. doi: 10.1002/jgre.20046.
A. Spiga, D. Banfield, N. A. Teanby, F. Forget, A. Lucas, B. Kenda, J. A. Rodriguez Manfredi, R. Widmer-Schnidrig,
N. Murdoch, M. T. Lemmon, R. F. Garcia, L. Martire, Ö. Karatekin, S. Le Maistre, B. Van Hove, V. Dehant,
P. Lognonné, N. Mueller, R. Lorenz, D. Mimoun, S. Rodriguez, É. Beucler, I. Daubar, M. P. Golombek, T. Bertrand,
Y. Nishikawa, E. Millour, L. Rolland, Q. Brissaud, T. Kawamura, A. Mocquet, R. Martin, J. Clinton, É. Stutzmann,
T. Spohn, S. Smrekar, and W. B. Banerdt. Atmospheric Science with InSight. Space Science Reviews, 214:109, 2018.
doi: 10.1007/s11214-018-0543-0.
P. M. Streeter, S. R. Lewis, M. R. Patell, J. A. Holmes, and D. M. Kass. Surface Warming During the 2018/MY 34
Mars Global Dust Storm. In Ninth International Conference on Mars 2019(LPIContrib.No.2089), page 6242, 2019.
A. Sánchez-Lavega, T. del Río-Gaztelurrutia, J. Hernández-Bernal, and M. Delcroix. The onset and growth of the 2018
martian global dust storm. Geophysical Research Letters, 46(11):6101 -- 6108, 2019. doi: 10.1029/2019GL083207.
URL https://agupubs.onlinelibrary.wiley.com/doi/abs/10.1029/2019GL083207.
M. Vincendon, J. Audouard, F. Altieri, and A. Ody. Mars Express measurements of surface albedo changes over
2004-2010. Icarus, 251:145 -- 163, 2015. doi: 10.1016/j.icarus.2014.10.029.
27
A PREPRINT - JULY 19, 2019
C. Wang, F. Forget, T. Bertrand, A. Spiga, E. Millour, and T. Navarro. Parameterization of Rocket Dust Storms on
Mars in the LMD Martian GCM: Modeling Details and Validation. Journal of Geophysical Research (Planets), 123:
982 -- 1000, 2018. doi: 10.1002/2017JE005255.
R. W. Wilson and K. Hamilton. Comprehensive model simulation of thermal tides in the Martian atmosphere.
J. Atmos. Sci., 53:1290 -- 1326, 1996.
Paul Withers. Empirical Estimates of Martian Surface Pressure in Support of the Landing of Mars Science Laboratory.
Space Science Reviews, 170(1-4):837 -- 860, Sep 2012. doi: 10.1007/s11214-012-9876-2.
R. W. Zurek. Martian great dust storm, an update. Icarus, 50:288 -- 310, 1982.
Richard W. Zurek and Leonard J. Martin. Interannual variability of planet-encircling dust storms on Mars. J. Geo-
phys. Res., 98(E2):3247 -- 3259, 1993.
28
A PREPRINT - JULY 19, 2019
Figure 19: Different versions of the gridded maps of 9.3 µm absorption column dust optical depth for SOY 400,
Ls ∼ 196◦, in the growing phase of the Global Dust Event of MY 34. V2.2 is the top left map, v2.3 is the top right one,
v2.4 is the bottom left one, and the reference v2.5 is the bottom right one. The spatial resolution of all gridded maps is
6◦ longitude × 5◦ latitude.
29
A PREPRINT - JULY 19, 2019
Figure 20: Time series of equivalent visible column dust optical depth calculated from the 9.3 µm absorption CDOD
normalized to 610 Pa, extracted from different versions of the gridded maps in an area around Gale crater, compared to
the time series of visible column optical depth measured by MastCAM aboard NASA's "Curiosity" rover (black line).
Curiosity observations [Guzevich et al., 2018, see main text] have been daily averaged and normalized to 610 Pa (using
the surface pressure from the Mars Climate Database pres0 routine). Both time series are shown between Sol-of-Year
355 and 500, i.e. Ls ∼ [170◦, 260◦]. We used a factor of 2.6 to convert 9.3 µm absorption CDOD into equivalent
visible ones. Data from gridded maps are averaged in longitude=[123◦E, 153◦E], latitude=[15◦S, 10◦N] centered
around Curiosity landing site at longitude 137.4◦E and latitude 4.6◦S.
30
|
1105.0045 | 1 | 1105 | 2011-04-30T04:41:14 | The Photoevaporative Wind from the Disk of TW Hya | [
"astro-ph.EP"
] | Photoevaporation driven by the central star is expected to be a ubiquitous and important mechanism to disperse the circumstellar dust and gas from which planets form. Here, we present a detailed study of the circumstellar disk surrounding the nearby star TW Hya and provide observational constraints to its photoevaporative wind. Our new high-resolution (R ~ 30,000) mid-infrared spectroscopy in the [Ne II] 12.81 {\mu}m line confirms that this gas diagnostic traces the unbound wind component within 10AU from the star. From the blueshift and asymmetry in the line profile, we estimate that most (>80%) of the [Ne II] emission arises from disk radii where the midplane is optically thick to the redshifted outflowing gas, meaning beyond the 1 or 4AU dust rim inferred from other observations. We re-analyze high-resolution (R ~ 48, 000) archival optical spectra searching for additional transitions that may trace the photoevaporative flow. Unlike the [Ne II] line, optical forbidden lines from OI, SII, and MgI are centered at the stellar velocity and have symmetric profiles. The only way these lines could trace the photoevaporative flow is if they arise from a disk region physically distinct from that traced by the [Ne II] line, specifically from within the optically thin dust gap. However, the small (~10 km/s) FWHM of these lines suggest that most of the emitting gas traced at optical wavelengths is bound to the system rather than unbound. We discuss the implications of our results for a planet-induced versus a photoevaporation-induced gap. | astro-ph.EP | astro-ph |
The Photoevaporative Wind from the Disk of TW Hya
Lunar and Planetary Laboratory, The University of Arizona, Tucson, AZ 85721, USA
I. Pascucci1
[email protected]
M. Sterzik
European Southern Observatory, Casilla 19001, Santiago 19, Chile
Department of Physics & Astronomy, University of Leicester, University Road, Leicester,
R. D. Alexander
LE1 7RH, UK
S. H. P. Alencar
Departamento de Fisica - ICEx UFMG, MG, Brazil
SETI Institute, 189 Bernardo Ave., Mountain View, CA 94043, USA
U. Gorti2
D. Hollenbach
SETI Institute, 189 Bernardo Ave., Mountain View, CA 94043, USA
J. Owen
Institute of Astronomy, Madingley Road, Cambridge CB3 0HA
B. Ercolano
University Observatory Munich, D-81679, Munich, Germany
1Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218
2NASA Ames Research Center, Moffett Field, CA 94035, USA
-- 2 --
S. Edwards
Astronomy Department, Smith College, Northampton, MA 01063 USA
Received
;
accepted
⋆Based on observations made with VISIR on the UT3/Melipal ESO Telescope at Paranal
under programme ID 084.C-0088(A)
-- 3 --
ABSTRACT
Photoevaporation driven by the central star is expected to be a ubiquitous
and important mechanism to disperse the circumstellar dust and gas from which
planets form. Here, we present a detailed study of the circumstellar disk sur-
rounding the nearby star TW Hya and provide observational constraints to its
photoevaporative wind. Our new high-resolution (R ∼ 30, 000) mid-infrared
spectroscopy in the [Ne ii] 12.81 µm line confirms that this gas diagnostic traces
the unbound wind component within 10 AU from the star. From the blueshift
and asymmetry in the line profile, we estimate that most (>80%) of the [Ne ii]
emission arises from disk radii where the midplane is optically thick to the red-
shifted outflowing gas, meaning beyond the 1 or 4 AU dust rim inferred from
other observations. We re-analyze high-resolution (R ∼ 48, 000) archival optical
spectra searching for additional transitions that may trace the photoevaporative
flow. Unlike the [Ne ii] line, optical forbidden lines from OI, SII, and MgI are
centered at the stellar velocity and have symmetric profiles. The only way these
lines could trace the photoevaporative flow is if they arise from a disk region
physically distinct from that traced by the [Ne ii] line, specifically from within
the optically thin dust gap. However, the small (∼10 km/s) FWHM of these lines
suggest that most of the emitting gas traced at optical wavelengths is bound to
the system rather than unbound. We discuss the implications of our results for
a planet-induced versus a photoevaporation-induced gap.
Subject headings: accretion, accretion disks -- infrared: stars -- planetary systems:
protoplanetary disks -- stars: individual (TW Hya)
-- 4 --
1.
Introduction
One of the yet unsolved questions in planet formation is how young stars lose their
disks. Planet formation certainly contributes to the clearing of primordial disk material. Its
first steps, the coagulation of sub-micron sized grains and their subsequent settling to the
disk midplane, have been identified toward many protoplanetary disks (e.g., Natta et al.
2007 for a review). However, planet formation is likely not the major disk dispersal
mechanism, as evinced by the small total mass in the planets (both in our solar system
and in other planetary systems) when compared to the mass of protoplanetary disks
(e.g., Hayashi et al. 1985). Furthermore, a disk dispersal mechanism based only on planet
formation would require shorter dispersal timescales in high metallicity environments
(Ercolano & Clarke 2010), which is not supported by recent observations of disk fractions
in low metallicity clusters of the extreme outer Galaxy (Yasui et al. 2009).
Current models of protoplanetary disk evolution suggest that viscous evolution
(accretion of gas onto the central star) and photoevaporation driven by the central star
(heating of disk gas to thermal escape velocities) are the main disk dispersal mechanisms
(e.g., Hollenbach et al. 2000; Gorti et al. 2009; Owen et al. 2010). Indeed, there is abundant
observational evidence that young stars are accreting disk gas:
from the optical and
ultraviolet excess continuum emission arising from the accretion shock (e.g., Calvet et al.
2000) to the broad permitted emission lines from infalling gas (e.g., Muzerolle et al. 1998).
Measurements of mass accretion rates using these diagnostics are now available for hundreds
of stars in nearby star-forming regions and qualitatively follow the time evolution predicted
by these models (Hartmann et al. 1998; Sicilia-Aguilar et al. 2010). However, at least two
other observables are not reproduced by models of viscously evolving disks. First, the disk
dispersal timescale, which is measured to be of just a few Myr, is predicted to be too long
using the viscosity inferred from mass accretion rate measurements. Both optically thick
-- 5 --
dust disks as well as gaseous disks should be numerous and detectable around stars that are
∼10 Myr-old, which contrasts with observations (e.g., Pascucci & Tachibana 2010). Second,
the transition from disk-bearing to disk-less stars is observed to be quick (just ∼105 years),
too quick to be explained by viscous evolution alone (see e.g., Armitage 2010 for a review).
These observables led theorists to propose photoevaporation driven by the central
star as the next ubiquitous and most relevant mechanism to disperse protoplanetary
disks (e.g., Hollenbach et al. 2000; Dullemond et al. 2007). In the first model combining
the effect of accretion and photoevaporation (Clarke et al. 2001), extreme UV photons
(EUV, 13.6 eV < hν ≤ 100 eV) from the central star heat and ionize the disk surface.
Beyond the radius at which the sound speed of the gas equals the local Keplerian orbital
speed (gravitational radius, rg), the hot gas becomes unbound and a photoevaporative
wind is established. When the disk accretion rate through rg drops below the wind loss
rate, photoevaporation limits the supply of gas to the inner disk, which drains onto the
star on the local viscous timescale -- of order 105 years. Many theoretical developments
have been made since this first model. First, Liffman (2003), Font et al. (2004), and
Adams et al. (2004) have included the effects of angular momentum support and showed
that photoevaporation occurs within rg and happens mostly outside ∼0.15 rg (hereafter
critical radius). Second, Alexander et al. (2006a,b) showed that the direct EUV irradiation
of the disk rim can drastically reduce the disk lifetime once the inner disk is cleared out.
Third, Ercolano et al. (2008), and Gorti & Hollenbach (2009) have recognized that stellar
Xrays and FUV photons can launch winds from denser disk regions and thus dramatically
increase the wind mass-loss rate produced by EUV photons alone.
Recently, several papers have discussed whether other disk properties, beyond disk
lifetimes, are consistent with photoevaporation models. There is yet no consensus:
Kim et al. (2009) include photoevaporation as a possible explanation for some of the
-- 6 --
transition disks in their sample; Currie et al. (2009) argue that the large number of evolved
dust disks at 5 Myr is inconsistent with photoevaporation; Cieza et al. (2008) find support
for photoevaporation based on the lack of optically thin inner disks with detectable dust
outer disks. All these studies and similar ones relied on comparing model predictions with
the properties of dust disks. A more direct way to test photoevaporation is to observe the
gas disk component, especially that component that is not gravitationally bound and is
part of the photoevaporating flow.
Following this approach, Pascucci & Sterzik (2009) have shown that the [Ne ii] line
at 12.81 µm is a good diagnostic for the photoevaporative wind in some systems that
have transition-like spectral energy distributions (SEDs). Such systems display small
near-infrared excess emission but large mid- to far-infrared emission and are thus thought
to be in the process of clearing out their primordial disk material. The most convincing case
for photoevaporation is that for the almost face-on transition disk around the 10 Myr-old
star TW Hya. The measured blueshift of ∼6 km/s demonstrates that there is unbound gas
leaving the disk and moving toward the observer (Pascucci & Sterzik 2009). If the wind is
fully ionized, the measured [Ne ii] flux translates into a mass-loss rate of ∼ 10−10M⊙/yr.
However, if the wind is mostly neutral, as predicted by models of Xray photoevaporation1,
then the [Ne ii] emission is a trace component of the wind and the observed line strength
and profile imply a much larger wind rate up to ∼ 10−8M⊙/yr (Ercolano & Owen 2010).
We also note that this relatively old star is still accreting disk gas at a rate that is about
an oder of magnitude lower than that of classical T Tauri stars (∼10−9 M⊙/yr, Muzerolle
et al. 2000 and Alencar & Batalha 2002).
1Gorti & Hollenbach (2009) point out that Xrays alone cannot drive a strong wind but
it is rather the combination of FUV photons (which dissociate the molecules and thereby
reduce the ability of the disk gas to cool) and Xrays (which heat the gas) that drive the wind
-- 7 --
Being at only 51 pc (Mamajek 2005) and clearly presenting evidence for photoevap-
oration, TW Hya is the ideal target to provide additional observational constraints to
models of disk photoevaporation and to help determine wind mass-loss rates. In this paper
we present new observations of the [Ne ii] 12.81 µm line toward TW Hya and employ the
spectroastrometry technique to estimate the radii traced by the ionized wind component.
We also analyze archival optical spectra to search for the neutral disk wind component
predicted by Xray-driven photoevaporation models. We discuss the implications of our
results on the gas distribution in the inner disk of TW Hya and speculate on how that
structure might link to planet formation and photoevaporation.
2. Photoevaporative disk wind models
Before presenting our observations and analyzing the infrared and optical spectra, we
wish to briefly summarize the main features of photoevaporative disk wind models and
point out their major differences. Our goal is to clarify what observables can be used to
characterize the structure and extension of the wind based on model predictions.
There are basically three main sources of ionization and heating of the disk surface:
Xrays (0.1 keV < hν < 10 keV), EUV (13.6 eV < hν < 0.1 keV), and FUV (6 eV < hν <
13.6 eV) photons. The first and simplest models assumed that photoevaporation is driven
solely by stellar EUV photons. These photons ionize the hydrogen in the very upper layers
of the disk and heat it to a temperature of ∼104 K, independent of radius, thus creating a
kind of coronal H II region (Hollenbach et al. 1994, Alexander et al. 2006a,b). This fully
ionized layer will produce forbidden lines at infrared and optical wavelengths whose relative
intensities depend on the intensity and the spectral energy distribution of the EUV field
impinging on the disk (Hollenbach & Gorti 2009), which is not constrained observationally.
Assuming plausible stellar ionizing fluxes of ∼ 1041 photons/s and a solar abundance of the
-- 8 --
elements, EUV-only models predict a relatively strong [Ne ii] emission line at 12.81 µm
(Alexander 2008, Hollenbach & Gorti 2009) and similarly strong [S II] lines at 6731 and
6716 A and [N II] lines at 6583 A (Font et al. 2004). Because the fraction of neutral gas
is very low in the EUV layer, forbidden lines from atomic species, such as the [O I] line
at 6300 A, are expected to be weak (Font et al. 2004, Hollenbach & Gorti 2009). The
most recent developments of the EUV-only models include a detailed treatment of the flow
structure and predict line profiles from the wind in addition to line intensities (Alexander
2008). As we will show in Sect. 4.4.1, these simple models can also produce channel maps
at specific emission lines that can be used to constrain the spatial distribution of the wind.
More comprehensive models, such as those developed by Ercolano et al. and Gorti &
Hollenbach, include other sources of disk heating and ionization in addition to stellar EUV
photons. Specifically, the most recent development of the Ercolano et al. models includes
Xray and EUV heating and ionization of the atomic disk gas and a 2D hydrodynamic
calculation wich allows predicting line intensities and profiles from a variety of forbidden
and semi-permitted transitions in the wind (Ercolano & Owen 2010). Hereafter, we will call
these models X-EUV models. Gorti & Hollenbach (2009) include all three main heating
and ionization sources and a detailed treatment of molecular cooling in their thermal
balance calculations. For this more complex model, which we will call XE-FUV model,
hydrodynamical calculations have not been perfomed yet hence we will limit the comparison
between model and observations to line intensities. The X-EUV and XE-FUV models
further differ in their assumptions of the stellar Xray and EUV spectra and therefore also
in the predicted mass-flow rates. Regardless of these differences, the inclusion of Xray
and FUV heating results in a predominantly denser, but cooler (hence mainly neutral),
photoevaporative wind compared to the EUV-only models and the mass loss rates are
higher by one (XE-FUV) or two (X-EUV) orders of magnitude. Depending on the star/disk
parameters, the [OI] emission line at 6300 A, whose low-velocity component is always
-- 9 --
detected toward accreting T Tauri stars (Hartigan et al. 1995), may trace the neutral
component of the disk wind (Ercolano & Owen 2010). Other forbidden lines that are
expected to be strong in the Xray ionized wind are the [Ne ii] line at 12.81 µm (Hollenbach
& Gorti 2009, Ercolano & Owen 2010) and the [S II] lines at 6731 and 6716 A (Ercolano &
Owen 2010).
We will now analyze our mid-infrared and optical spectra focusing on gas lines that
could trace the ionized and neutral components of the disk wind.
3. New mid-infrared observations and data reduction
Our first VLT/VISIR observations have shown that the [Ne ii] 12.81µm emission from
TW Hya traces unbound gas moving toward the observer, likely in a photoevaporative
disk wind (Pascucci & Sterzik 2009). We performed additional long-slit high-resolution
spectroscopy with VISIR (Lagage et al. 2004) on 23-24 February 2010 attempting to detect
the photoevaporative flow at a resolution better than the instrument resolving power
(∼ 400 mas or 20 AU at the distance of TW Hya). To reach this goal we have applied the
so-called spectro-astrometric technique (e.g., Whelan & Garcia 2008) and acquired spectra
at four different slit orientations (position angle PA = 0, 270, 333, 63◦) interleaved by four
antiparallel orientations, i.e. where the slit is rotated by 180◦ (position angle PA = 180, 90,
153, 243◦), see Table 1. PAs 0◦ and 180◦ correspond to N-S slit orientations on the sky,
270◦ and 90◦ to W-E while the other two couples of PAs were chosen to cover the PA of
the disk of TW Hya (i.e. the angle between N and the major axis of the disk measured E
of N, see Table 2) and a direction perpendicular to it. As in past observations, we used a
slit width of 0.′′4 which provides a resolution of R∼30,000 or ∼10 km/s in velocity scale at
∼12.8µm (Kaufl 2006; Pascucci & Sterzik 2009).
-- 10 --
Before each spectroscopic integration we acquired TW Hya with the PAH2-NEII filter
to accurately position the source in the slit. Fig. 1 shows the difference in pixels between
the location of the target and the slit center in cross-dispersion direction, i.e. perpendicular
to the slit length. The FWHM of the slit (∼ 3 pixels) is also overplotted. Except for
PA=63◦, TW Hya was positioned within half of the slit FWHM from the slit center. We
note, however, that for this specific PA the emission is also the faintest, hence the error in
the source location is the largest (dashed line in Fig. 1). The source faintness could be due
to flux loss caused by the imperfect centering.
We applied the standard chopping/nodding technique with a throw of 8′′ along the slit
to suppress the mid-infrared background. Based on our previous experience, we set the
on-source exposure time to 1 h for each slit orientation. For all PAs in the first night and
for the first PA in the second night we also observed a standard star immediately before
or after the TW Hya spectroscopic exposure and with the same slit PA to correct the
spectra for telluric absorption and to obtain an absolute flux calibration. A summary of the
observations is presented in Table 1.
Each spectroscopic exposure has been reduced as described in Pascucci & Sterzik
(2009) using the VISIR pipeline version 3.2.2 (Lundin 2008). In brief, images are first
corrected by fluctuations in the background using the off-source chop and nod exposures.
In parallel, a reference frame of the infrared background is also created. Next, images are
corrected for the optical distortion, and are shifted and added to form a final combined
image from which the spectrum is extracted following the optimal extraction method by
Horne (1986). Finally, the spectrum is wavelength calibrated by cross correlating the
background frame spectrum to a synthetic model spectrum of the atmosphere in the
observed wavelength range (note that the resulting wavelengths are in vacuum).
To remove telluric features we have used ATRAN model atmospheres for different
-- 11 --
elevations and amounts of precipitable water as follows. We normalized each spectrum of
TW Hya to the continuum, such that the continuum is 1, and scaled the model atmospheres
to minimize (in a χ2 statistics) the difference between the observations and the model in
regions where there are strong atmospheric absorption features (the strongest one is from
CO2 around 12.812 µm). During the minimization we took into account the flux uncertainty
at each wavelength and also allowed for a small shift of ±1 pixel in wavelength to find
the best scaled/shifted atmospheric model that matches the continuum of TW Hya. Since
in the first night we also acquired standard stars immediately before or after each PA of
the TW Hya spectra, we could verify that the approach described above gives identical
results (within the flux uncertainties) to the more traditional approach of dividing each
science spectrum by that of a flux calibrated standard star spectrum. Fig. 2 presents the
five TW Hya spectra that could be flux calibrated while Table 3 summarizes the main
properties of the observed [Ne ii] lines for all PAs.
3.1. Accuracy of the VISIR wavelength solution
Because one of our goals is to measure small shifts in the peak of the [Ne ii] line, we
performed additional tests to verify the instrument wavelength solution and investigate
the relative and absolute wavelength accuracy of VISIR in the high-resolution long-slit
mode. First of all, we reduced and analyzed 8 additional archival spectra from 3 different
infrared-bright K4-6 III standard stars (HD 136422, HD 139127, HD 149447) observed in
the same setting as TW Hya. After applying the proper heliocentric correction and shifting
the spectra in the stellocentric frame, we compared the observed peak position of two
photospheric absorption lines to those predicted by the MARCS model atmosphere of a
K5 III star (HR 6705; the spectrum was kindly provided by L. Decin). While differences
in velocities are confined within ±2 km/s it is apparent that one of the lines is typically
-- 12 --
redshifted while the other is typically blueshifted. To investigate whether there could be
an issue in the wavelength dispersion solution and whether errors could be larger at other
wavelengths, we turned to spectra of Titan which were acquired with the same setting as
TW Hya (VLT/VISIR PID 083.C-0883). Titan spectra present many unresolved lines from
C2H2 and C2H6 molecules in this setting. In case of large drifts in wavelength, we might see
the FWHM of unresolved lines to change across the spectrum. We measured the FWHMs
of 7 unresolved emission lines in two spectra of Titan covering the full VISIR wavelength
range (3 before and 4 after the [Ne ii] line in TW Hya) and found no such trend, rather we
report a narrow range of FWHMs with a mean value of 10 km/s, the spectral resolution
of VISIR in this setting. In addition, we compared the location of Titan lines with the
rest wavelengths reported in the HITRAN database (Rothman & Gordon 2009). In the
setting covered with VISIR there are seven transitions from the C2H6 molecule, 2 of which
are single and can be used to measure the relative centroids to other lines through the
spectrum. Fig. 3 shows that the relative uncertainty between the lines is 1 km/s, one-tenth
of the spectral resolution of VISIR in this mode. However, as hinted by the test on the
standard stars and as we will see on the TW Hya data (Sect. 4), the absolute accuracy
reachable with VISIR is likely a factor of 2 higher (2 km/s), due to uncertainties in placing
the source in the center of the slit.
4. Constraints on the wind structure from ionized and neutral species
In the following we analyze our new mid-infrared high-resolution spectra with special
emphasis on characterzing the [Ne ii] peak velocities and the overall line profile. We also
re-analyze already published high-resolution optical spectra of TW Hya to search for
additional transitions from neutral and ionized species that may trace the photoevaporative
flow. The optical spectra were obtained between 1998 and 2000 with the FEROS
-- 13 --
echelle spectrograph on the 1.5 m ESO telescope on La Silla at a resolution of ∼ 6 km/s
(R ∼48,000). Alencar & Batalha (2002) presented an extensive analysis of the optical
variability of TW Hya based on this dataset. We refer to their paper for details on the data
reduction.
4.1.
[Ne ii] line fluxes and peak velocities
Spitzer/IRS spectra of TW Hya taken in three different years show variations of
up to ∼30% in the continuum emission, as well as in the [Ne ii] flux and equivalent
width (Najita et al. 2010). These differences may arise from pointing errors and/or true
source variability. The VISIR spectra we acquired in this observational campaign provide
additional epochs to further investigate continuum and line variability on the timescale of
years as well as on the much shorter timescale of hours.
For each VISIR spectrum we compute the location of [Ne ii] peak emission in
the stellocentric frame (after assuming a stellar radial velocity of 12.2±0.5 km/s,
Weintraub et al. 2000), the EW, and when standard stars were acquired close in time to the
TW Hya spectroscopic exposures, also the line flux and continuum near the line (essentially
for each PA in the first night and for the first PA in the second night). Peak positions
and [Ne ii] line fluxes are calculated from gaussian fits to the observed line profiles, while
EWs are computed by integrating the line flux within ±4σ of the fitted gaussian. Table 3
summarizes the main results and Fig. 2 shows line profiles for the five flux calibrated
spectra.
To test the hypothesis of variations due to calibration uncertainties, we search for
correlations between the main properties of the [Ne ii] line and the adjacent continuum
with the airmass and the location of the source in the slit. We find an anti-correlation
-- 14 --
between the line flux and continuum and the source airmass (lower airmass→higher flux
and continuum, see Fig. 4) suggesting that most variations in these quantities are due to
flux calibration uncertainties. We do not find any correlation between these quantities and
the location of the source in the slit. However, we note that the displacement of the source
with respect of the center of the slit might be the cause for the lowest EW and the largest
blueshift in the [Ne ii] line for PA=63◦ (see Fig. 1 and Fig. 4). Unfortunately, we do not
have a standard star to flux calibrate the TW Hya spectrum at this PA but our expectation
is of a reduced line flux and continuum. We also note that the highest [Ne ii] flux we
measure (for PA=180◦, smallest airmass) is as high as that measured by Spitzer/IRS within
the estimated errorbars. This demonstrates that the small VISIR slit recovers all the [Ne ii]
emission from the disk of TW Hya implying that the emission is confined within ±10 AU
from the star.
[Ne ii] line EWs are less affected by differences in the airmass: for the 5 exposures at
airmass < 1.3 (excluding PA=63◦) the mean EW is 46±2A. This value is slightly larger than
the [Ne ii] EW from our 2008 high-resolution spectrum (Pascucci & Sterzik 2009), the same
as the 2006 EW measured with Spitzer/IRS (Najita et al. 2010), and smaller by about 20 A
than the 2004 and 2008 EWs measured with Spitzer/IRS and the 2007 EW measured with
Gemini/Michelle by Herczeg et al. (2007). Although the sampling is sparse, the [Ne ii] EW
essentially oscillates around two values (∼45 and 65 A) which suggests intrinsic variability
in the [Ne ii] emission or nearby continuum on a timescale of ∼1-2 years.
We also confirm the blueshift in the [Ne ii] line with respect to the radial velocity of
TW Hya (see Fig. 4 second panel). The mean blueshift is -5±1 km/s in agreement with
our previously reported blueshift (Pascucci & Sterzik 2009). The peak position of the
[Ne ii] line is not correlated with the airmass at which the source was observed, nor with
the location of the source in the slit (but see previous note for PA=63◦). The FWHM of
-- 15 --
the [Ne ii] line presents the least variations with PA, is not correlated with the EW, peak
position, nor with the airmass: the mean value and standard deviation for the FWHM
are 16.3±0.9 km/s, or an intrinsic FWHM of 12.9±0.9 km/s after deconvolution with the
instrumental FWHM of VISIR of 10 km/s.
4.2. Forbidden and semi-permitted optical lines
We search the FEROS optical spectra for all the lines predicted by Font et al. (2004),
Hollenbach & Gorti (2009), and Ercolano & Owen (2010) to have luminosities of a few
10−7 L⊙ or higher, thus comprising forbidden and semi-permitted transitions from OI,
OII, MgI, NII, and SII. We report equivalent widths and upper limits in Table 4 for
representative transitions. To search for weak lines and properly measure EWs we had to
correct for the numerous photospheric absorption lines present in the optical spectra of
TW Hya. We followed the approach described by Alencar & Batalha (2002) and divided
the continuum normalized spectra of TW Hya with the artificially veiled and continuum
normalized spectrum of GJ 1172, a K7 V star with similar v sini as TW Hya (see also
Hartigan et al. 1989 for the procedure and definition of veiling). In addition, we combine
all observed spectra in a median high signal-to-noise spectrum normalized to the continuum
level.
Optical spectra from T Tauri stars are known to have strong forbidden emission lines
from [OI] at 6300, 6364, and 5577 A (Hartigan et al. 1995). TW Hya is no exception and
these transitions are easily identified in all individual spectra. The strongest of all is the
[OI] line at 6300 A with a median EW of 0.47 A, which was found not to vary significantly
with time (Alencar & Batalha 2002). Following Hartigan et al. (1995) and assuming a
mean R magnitude of 10.2, we convert this EW into a line luminosity of 10−5 L⊙. The
optical spectrum of TW Hya is known to vary, the most detailed V magnitude monitoring
-- 16 --
reports night-to-night variations of ∆V ≃ 0.35 (seasonal changes have smaller amplitudes,
Rucinski et al. 2008). The same variations in R convert into a [OI] luminosity range of
0.8-1.6×10−5 L⊙. The 6364 A [OI] line is a factor of ∼3 weaker than the [OI] 6300 A, as
expected from the ratio of the Einstein coefficients, while the 5577 A [OI] line is a factor of
∼7 weaker on average (see Table 4). These lines are also spectrally resolved with FEROS,
the mean (and median) FWHM of the [O i] 6300 A is 11.5±0.4 km/s or an intrinsic line
width of ∼10 km/s after deconvolution with the 6 km/s instrumental FWHM. The other
weaker [OI] lines have similar FWHMs. As already reported in Alencar & Batalha (2002)
the [O i] 6300A line presents only a narrow component, is centered at the stellar rest frame,
and there is no indication of a blueshifted high-velocity component from jets/outflows. We
use 40 spectra from Alencar & Batalha (2002) to further investigate whether there is a
blueshift of just a few km/s in the [O i] peak emission2. To do that we rely not only on
the absolute wavelength calibration performed previously but also compare the location
of the [O i] 6300 A line with respect to other known photospheric lines. Fig. 5 shows the
distribution of velocities for the [O i] 6300 A line in the stellocentric frame measured via
the photospheric CaI 6439 A line (black dashed line). This histogram peaks at negative
velocities but the median difference velocity is only -0.1 km/s with a standard deviation as
large as 0.15 km/s. While this tiny blueshift might indicate a flow, we show in the same
figure that it is almost certainly related to small differences in the dispersion solution
2We excluded the 10 spectra obtained by Alencar & Batalha (2002) in December 1998
because the terrestrial night-sky emission line at [OI] 6300A falls at +12 km/s in the stel-
locentric frame thus introducing a clear bump on the red side of the [O i] line from TW Hya.
For the other 40 spectra obtained in April and February the terrestrial [O i] line is blueshifted
by 23 km/s and is within 1 km/s from the [O i] line of TW Hya respectively. Thus, the [O i]
line profiles from TW Hya are not affected by the night-sky emission in these 40 spectra
-- 17 --
between the two different echelle orders in which the [O i] 6300 A and the CaI 6363 A fall.
To demonstrate this we overplot the difference between the [O i] 6300 A and the [O i] 6364 A
which falls in the same order as the CaI photospheric line and should trace the same disk
emission as the [O i] 6300 A line. Fig. 5 clearly shows that the two velocity distributions
are identical, the median difference between the [O i] 6300 A and the [O i] 6364 A being
-0.1 km/s and standard deviation being 0.16 km/s. These tests demonstrate that the [OI]
optical lines are not blueshifted.
In addition to the oxygen lines, we also detect the [S II] line at 4069 A and the Mg I]
line at 4571 A with luminosities that are about ∼10 times lower than the [OI] 6300 A
luminosity (Table 4). The [S II] transition at 4076 A comes from very similar upper energies
(∼3 eV) as the 4069 A transition but the Einstein coefficient is 2.5 times smaller, consistent
with our non-detection and 3σ upper limit. The [S II] at 4069 A and the Mg I] line at
4571 A peak at the stellar radial velocity as the [O I] lines. Their luminosities are similar
to those predicted by the X-EUV primordial disk model with Log(Lx)=29.3 (Ercolano &
Owen 2010). However, we note that this Xray luminosity is low for TW Hya, about an order
of magnitude lower than what has been estimated by e.g. Kastner et al. (1999). In the
framework of photoevaporative disk models, the most puzzling result is the non-detection
of the other two [S II] transitions at 6716 and 6731 A with stringent upper limits pointing
to fluxes ∼50 times lower than the [O I] 6300 A line. As mentioned in Sect. 2, both EUV
and Xray irradiated disks should produce a strong [S II] doublet for gas at low density
and flowing, similar in strength or just a factor of a few weaker than the [OI] 6300 A line.
While the [SII] 6716/6731 doublet should trace more extended low density gas than other
forbidden lines due to its much lower critical density3, it is unlikely that the FEROS spectra
3according to EUV models the radial extension is a factor of 2 greater than the [Ne ii]
emission (out to about ±20 AU from the star) and less concentrated to the center
-- 18 --
lost most of the [SII] 6716/6731 flux since the fiber diameter projected on the sky is 2.7′′
and covers radii out to ∼70 AU at the distance of TW Hya. Thus, the non-detection of
these transitions points to specific conditions in the disk emitting region, specifically to
dense gas, as we shall discuss later.
Finally, the centrally peaked profiles and the FWHM of the optical lines can help
constrain the extension of the emitting region. The [OI] 6300 A line has an intrinsic width
of 10 km/s, the [SII] and MgI] lines are similarly narrow ∼10-12 km/s, hence likely trace
similar disk radii. For a given ring of gas in Keplerian rotation extending out to a radius
ro from the star the absence of a double peak implies: 2q( GM⋆
ro >0.4-1 AU for the mass and disk inclinations reported for TW Hya (Table 2). Again for
ro )sin(i) < ∆vFEROS or
gas in Keplerian rotation, the inner ring radius ri determines the maximum velocity in the
line profile. Because the observed [OI] 6300 A profiles extend out to ±10 km/s (see Sect. 4.3)
the inner radius of the gas ring should be <0.2 AU and can be as close to the star as 0.06 AU
for a 4◦ disk inclination. This inner disk gas is likely not photoevaporating. However, if
most of the emission is extended beyond about 1 AU and gas is photoevaporating, detailed
disk models are necessary to compute line profiles and estimate the extension of the emitting
region (see Sect. 5.2).
4.3. Comparison of [Ne II] and [O I] line profiles
Our first high-resolution VISIR spectrum of TW Hya showed a hint for an asymmetry
in the [Ne ii] line profile (see Fig. 4 from Pascucci & Sterzik 2009), with possibly more
emission on the blue side than on the red side of the line. A flux enhancement on the blue
side was also pointed out by Herczeg et al. (2007) in their lower signal-to-noise spectrum.
Such an enhancement is expected for face-on disks with an optically thick midplane blocking
the view of the redshifted outflowing gas (Alexander 2008; Ercolano & Owen 2010). We use
-- 19 --
the 8 additional spectroscopic exposures acquired in this observational campaign to further
investigate the [Ne ii] 12.81 µm profile. To check for asymmetries we apply the following
method: We produce for each line a mirror spectrum (in velocity scale), then we apply small
shifts in velocities to the mirror spectrum such that the upper half of the lines matches
best, finally we subtract the shifted-mirror and original spectra. This method is especially
sensitive to detect any blue or red excess emission in the wings of the line compared to its
core emission and is more suited to lines that are poorly sampled in velocity than the line
bisector method (e.g., Gray 1980).
Fig. 6 summarizes our findings. Individual line profiles as well as the median of line
profiles (thick black line) show an excess emission on the blue side of the [Ne ii] line peaking
around -20 km/s and a typical shift of the peak centroid of -5 km/s4. The blue excess is also
visible as decreased emission on the red side of the line at ∼10 km/s (see upper and lower
panels of Fig. 6). We can confirm that this asymmetry is intrinsic to the source emission
and not an instrumental effect by applying the same method to spectrally unresolved
emission lines from Titan (see Fig. 7).
We then apply the same method of mirrored spectra to simulated [Ne ii] line profiles
from photoevaporative disk winds. As explained in Sect. 2, such profiles are available for
the EUV-only (Alexander 2008) and the X-EUV (Ercolano & Owen 2010) models. These
models use input parameters such as stellar mass, Xray and EUV luminosities that are
similar to measured values for the TW Hya system but were not specifically tuened for
4The two most different curves are for PA=270 and 333◦. The line profile from the first PA
shows decreased emission (one pixel) at ∼20 km/s and a positive (one pixel) emission at ∼-
30 km/s. The subtraction of the mirror-shifted profile (having reduced emission at -30 km/s
and excess at 20 km/s) generates the pronounced peak at -30 km/s in the original-mirror
profile. The line for PA=333◦ is consistent with a symmetric profile.
-- 20 --
TW Hya (in the case of the X-EUV we use the models with Log(LX)=30.3 which are
appropriate for TW Hya). In all cases the disk midplane is optically thick thus introducing
an asymmetry in the line profile which is, as expected, more pronounced the closest the
disk is to the face-on orientation (see Fig. 9). Applying the KS test to the difference
curves shows that the XEUV 10◦ disk inclination model has just a 6% probability that
the observed and model profiles are drawn from the same parent population hinting to a
difference. All other models (EUV with i =3, 10◦ and XEUV with i =3◦) have probabilities
as high as 40% meaning that they are all equally good matches to the data. The important
result from this analysis is that the [Ne ii] emission must arise from a region in the disk
that is optically thick to the redshifted outflowing gas. We shall discuss in Sect. 5 what this
implies for the extension of the ionized wind component of the photoveporative flow.
Finally, we also apply the method of mirrored spectra to the strongest of the optical
lines, the [OI] 6300 A line, to check for asymmetries in the line profile. Fig. 8 shows that,
unlike the [Ne ii] line, the [OI] profile is symmetric and the line is centered at the stellar
velocity. As we will discuss in more detail in Sect. 5, this profile could result from:
i)
pure Keplerian motion of gas bound to the disk (i.e. the [O i] is not tracing the neutral
component of the wind, it is not flowing); or ii) a photoevaporative flow from a disk region
that is optically thin to the redshifted outflowing gas. Examples of symmetric lines from
photoevaporative flows are presented in Ercolano & Owen (2010) for disks with inner holes
in their dust and gas distribution and emission confined within the hole. Because most of
the opacity is expected to be from dust (see Sect. 5) and dust extinction decreases with
wavelength, our line profile analysis demonstrates that if scenario ii) applies the neutral
[O i] gas does not trace the same radii as the [Ne ii] emitting gas.
-- 21 --
4.4. Spectroastrometric signal of a photoevaporative wind
We attempted for the first time spectroastrometry in the mid-infrared to improve upon
the spatial resolution achievable with the VLT/VISIR and resolve the innermost region of
the photoevaporative wind. Spectroastrometry relies on the fact that the measurement of
the centroid of a flux distribution can be determined to much better than a pixel precision
depending on the signal-to-noise of the observations. In the case of VISIR one pixel
corresponds to 127 mas meaning that we could investigate the [Ne ii] emission at scales
better than 6 AU at the distance of TW Hya. Asymmetries in the emission at a specific
wavelength appear as shifts in the centroid of the flux distribution. Using different slit
PAs enables to determine the orientations at which asymmetries are stronger and thus help
constraining the distribution of the source emission.
To retrieve the spectroastrometric signal from our VLT/VISIR data we extract spatial
profiles in each velocity channel and measure the spatial centroid of the [Ne ii] emission
in cross-dispersion direction with respect to the centroid of the emission in the continuum.
The continuum is measured as the mean of the continuum centroids on the red and on
the blue side of the [Ne ii] line. Because of the low S/N on the continuum we summed up
the spatial profiles over more than 10 pixels in velocity, we have tested that our results
do not vary if we change the size of the velocity bins or the location where we measure
the continuum. To determine the centroid of the spatial profiles we have tested Gaussian
and Lorentzian profiles as well as the weighted mean over the FWHM of VISIR in spatial
direction as proposed by Pontoppidan et al. (2008). We find that fitting the emission with
Lorentzian and Gaussian profiles provide the same spectroastrometric signal and a more
robust centroid than the weighted mean approach. As uncertainty on the centroid we take
the inverse of the S/N for each spatial profile where the noise is measured as the standard
deviation of the continuum outside the profile. This error reflects that profiles extracted
-- 22 --
at velocities sampling the wings of the line have lower S/N, hence the centroid is less well
determined. The centroid is also diluted by the continuum flux at each velocity to take into
account the different continuum-to-line ratio moving from the core to the wing of the line
(e.g., Whelan & Garcia 2008). Fig. 10 shows the centroids in pixels versus velocities in the
stellocentric frame for the parallel and anti-parallel slit orientations separately. This way
we can better search for any possible instrumental effects on the centroids because such
effects would appear as shifts of the same sign and intensity in the couple of parallel and
anti-parallel slits (Bailey 1998). The plot does not show any strong similar effect but small
amplitude modulations across the [Ne ii] line and nearby continuum. As we will see next
the centroids are measured to an accuracy of a tenth of a pixel close to the [Ne ii] peak
emission and typically to half of a pixel in the wings where the S/N of the flux distribution
in cross-dispersion direction is low5.
4.4.1. Comparison between observations and EUV-model predictions
The simple EUV-only disk model is computationally fast enough that we can create
channel maps for the [Ne ii] emission and compute spectroastrometric curves for the slit
orientations used in this observational campaign. The model is essentially identical to
the EUV-only model used by Font et al. (2004) and Alexander (2008) with a radially
continuous gas distribution. The only differences between our hydrodynamic model and
that of Alexander (2008) are the resolution of the model grid, and the location of the
lower boundary. As we wish to generate spatially-resolved emission maps, it is necessary
5We note that the low accuracy in the centroid for PA=63◦ is due to the lower flux in the
line and continuum likely caused by the fact that the source was not well centered in the slit
for this PA, see Fig. 1
-- 23 --
to use higher spatial resolution than that used by Alexander (2008). In addition, we now
assume that the underlying disk has a finite thickness, and place the base of the ionized
wind at z/r ≃ 0.15 (e.g., Hollenbach et al. 1994). This has only a small effect on the
predicted line profile, reducing the peak blue-shift by approximately 1 km/s. Our polar
[(r, θ)] computational grid covers the range r = [0.03rg, 10.0rg], θ = [0, 75◦], with grid cells
that are logarithmically spaced in r and linearly spaced in θ. We use Nr = 1113 cells in the
radial direction and Nθ = 250 cells in the polar direction. When the grid is expanded to
three dimensions (by assuming reflective symmetry about the disk mid-plane and azimuthal
symmetry around the polar axis), the resulting grid has 1113 × 600 × 1200 cells in the r, θ
and φ directions respectively.
We compute the line emission from each grid cell in the same manner as Alexander
(2008). However, following the results of Hollenbach & Gorti (2009) we adopt an ionization
fraction for Ne ii of 0.75 (rather than the somewhat arbitrary value of 0.33 used previously).
For comparison to the TW Hya system we adopt a stellar mass of M∗ = 0.7 M⊙ (thus
rg = 6.2 AU) and a disk inclination of i = 4◦ (see Table 2 for references). We set the ionizing
flux Φ by matching the predicted line luminosity to the observed value of 3.1 × 10−6L⊙:
we find that Φ = 7.5 × 1040 ionizing photons per second matches the observed line flux
to ∼ 1% accuracy. The integrated mass-loss rate from the disk is ∼ 1.5 × 10−10 M⊙. We
create an integrated line profile (identical to that of Alexander 2008) by summing the
contributions from every cell on the grid, and also generate spatially resolved emission
maps by de-projecting the three-dimensional grid on to the sky plane. This is done at
each (spectral) velocity to generate a three-dimensional data-cube. The model data cube
has 201x201 pixels which are 0.255 AU on a side (equivalent to 5 mas at the 51 pc distance
of TW Hya), and a velocity resolution of 0.1 km/s. We model the spectral resolution of
VISIR by convolving the data cube (in the velocity direction) with a Gaussian profile with
half-width σ = 5 km/s, to give a simulated spectral resolving power of R = 30, 000.
-- 24 --
Fig. 11 shows two simulated images of the [Ne ii] emission taken at a negative and
a positive velocity, the image orientation is as in Hughes et al. (2011) when N is up and
left is E in the figure. To simulate the spectroastrometric signal we apply a slit in the
x-axis equal to the VISIR slit width of 400 mas, collapse the flux in the x-axis and evaluate
the centroid of the spatial profile in the y-axis after convolution with the VISIR PSF
(FWHM of 300 mas) for each velocity channel. To simulate the rotation of the slit we
rotate the simulated images by the corresponding angle and then apply the slit as described
above. Fig. 12 shows the expected spectroastrometric signal as parallel minus anti-parallel
centroids in VISIR pixels for the photoevaporative disk model (solid lines) and for the wind
only, i.e. when Keplerian rotation is set artificially to zero (dashed lines). One sees that
for a slit orientation close to the disk major axis (PA=333◦) the spectroastrometric signal
is dominated by the Keplerian component of the gas, the asymmetry in the emission is
small but present due to the small but non-zero disk inclination. The positive displacement
at negative velocities can be understood by looking at Fig. 11: the emission at negative
velocities is stronger in the direction of PA=333◦ and slit orientations close to it. The wind
component dominates the signal for the slit orientation perpendicular to the disk major axis
(63◦ but it is prominent also for 270◦) because at that orientation the Keplerian component
is close to zero. The main results from the modeling are as follows: i) the expected shift in
centroid is small, about 0.2 pixels, and ii) the shift is primarily set by projection effects (disk
inclination) and the Keplerian component of the gas since the wind is more symmetrically
distributed around the star.
To compare model predictions with observations, we subtract the measured anti-parallel
centroid positions from the parallel positions (after interpolation over the same velocity
grid), see Fig. 13. For PAs 270 and 63◦, slits oriented close to the disk minor axis, the
spectroastrometric shift is less than 0.1 pixel as expected. For the other two slit orientations,
sampling the disk major axis, we do not detect the S-shape profile predicted for gas in
-- 25 --
Keplerian rotation. We see a hint of such a profile only for PA=333◦ negative velocities
which is however not present at the positive velocities. A smaller shift than predicted could
result from a [Ne ii] emission that is more symmetric than predicted, from a smaller disk
inclination, and/or from [Ne ii] emission more closely concentrated to the central star (in
this last case, the convolution with the VISIR PSF would smear out the contrast between
the bright and faint rims of the [Ne ii] emission). Further observations at even higher
spectral resolution and sensitivity, especially in the wings of the line, are necessary to
confirm these first results and to possibly identify the small spectroastrometric shift from
photoevaporating gas that is expected for slit orientations along the minor axis.
5. Physical implications on the disk wind
Before discussing the possible wind structure from the disk of TW Hya we wish to
briefly summarize the main observational results presented in the previous Sections.
• We confirm that the [Ne ii] emission line at 12.81 µm is blueshifted by ∼5 km/s with
respect to the stellar velocity. The line is relatively narrow (FWHM∼13 km/s) and
the profile is asymmetric, with more emission on the blue than on the red side.
• None of the forbidden and semi-permitted optical lines investigated here appear
blueshifted. On the contrary, these lines are centered in the stellocentric frame to
within about 0.15 km/s. The line profiles are symmetric and the similar FWHMs
(∼10 km/s) from different transitions point to optical lines tracing gas at similar radii.
• The very small or zero spectroastrometric shifts for the observed PAs suggest that the
[Ne ii] emission from the photoevaporative wind is predominantly spatially symmetric.
-- 26 --
What are these results telling us about the structure of the photoevaporative wind and
its extension?
5.1. The unbound disk component
The small blueshift in the [Ne ii] line is a clear tell-tale sign of photoevaporating gas
for a face-on configuration like that of TW Hya. Hence, we are sure that this line is probing
the ionized component of the disk wind. This component could be the dominant wind
component in the EUV case or a trace component of the mostly neutral wind in the Xray
case. Our observations constrain the radial extension of the ionized wind component via
the [Ne ii] line. By recovering all flux measured with Spitzer/IRS, we demonstrate that
the [Ne ii] emission arises within 10 AU from the star. We note that although the radial
extension of the mass loss is very different for pure EUV and X-EUV irradiation, with Xrays
producing 50% of the mass loss at tens of AU (Owen et al. 2010), the [Ne ii] luminosity
decreases steeply with radial distance from the star (as ∼ 1/r) hence the emission is
dominated by the flux from small radii. For the specific EUV and X-EUV models discussed
here 90% of the [Ne ii] emission arises within 12 AU and 15 AU respectively, which is
consistent with the 10 AU radial extension inferred observationally.
The blue excess in the wing of the [Ne ii] line adds a further constraint on the radii
traced by this wind component. We have shown in Sect. 4.3 that primordial disks (no gaps)
irradiated by EUV and Xrays can reproduce the [Ne ii] line profile and peak emission. Here
we ask the question: how much [Ne ii] emission could originate in an optically thin region
and still be consistent with the observed [Ne ii] profile? To simulate the [Ne ii] profile from
optically thin disk radii we take a gaussian centered at zero velocity and FWHM set by
thermal broadening for gas at ∼5,000 K (Xray heating) or ∼10,000 K (EUV heating) and
convolved with the FWHM of the VISIR spectrograph. We add as much as 10, 20, and 30%
-- 27 --
of the total emission to the [Ne ii] profile from a fully thick disk and compare the resulting
profile with the observed one. We find that already a 10% contribution from radii optically
thin to the [Ne ii] emission shifts the line peak emission redward to ∼-4 km/s, marginally
consistent with our measurements, and increases the symmetry of the line. This simple
approach shows that more than 80% of the [Ne ii] emission should come from a disk region
that is fully optically thick to the [Ne ii] line. What is then the source of opacity within
10 AU from the central star? This is a relevant question for transition-type sources like
TW Hya whose SEDs clearly point to opacity gaps in their disks.
First we look at the gas component and show that the [Ne ii] line is optically thin and
gas alone cannot be responsible for the blueshift and line asymmetry. Hollenbach & McKee
(1989) calculated that a column of Ne+ of a few times 1018 cm−2 (or a few 1022 H atoms for
a solar Ne/H abundance ratio) is necessary to produce an optical depth of 1 at the center of
the [Ne ii] line. This density has to be compared with < 1016 cm−2 Ne+ in the EUV layer
and < 1017 cm−2 in the Xray layer (Hollenbach & Gorti 2009; Glassgold et al. 2007), clearly
not sufficient to make the line optically thick. That the line is optically thin is also clear
from the observed line profile which does not show any absorption at the line center, where
the optical depth is the highest.
Sub-micron dust grains are the next major source of opacity. According to the disk
model of Calvet et al. (2002) the inner disk of TW Hya could have only a tiny amount
(0.5 lunar masses) of micron-sized grains out to ∼4 AU and thus would be optically thin
at visible and even more at infrared wavelengths out to that radial distance. However,
Ratzka et al. (2007) show that this model gives a factor of 2 too high mid-infrared visibilities
as well as the wrong wavelength dependence of the visibility. They argue that the dust
inner rim, hence the transition from thin to thick disk, must happen much closer in at about
0.5-0.8 AU. More recently, Akeson et al. (2011) also argue for the presence of optically
-- 28 --
thick dust within 4 AU. They propose a more complex structure with an optically thick
dust ring centered at 0.5 AU, a dust gap, and the full optically thick dust disk starting at
∼4 AU. Our simple approach of combining [Ne ii] profiles from optically thin and thick
disk radii suggests that more than 80% of the [Ne ii] emission should arise from disk radii
where the midplane is optically thick to the [Ne ii] emission. If the optically thin disk
extends out to 4 AU as in the model of Calvet et al. (2002), then the observed [Ne ii] profile
cannot be reproduced by the EUV-only and X-EUV models from Alexander (2008) and
Ercolano & Owen (2010) because in these models ∼50% of the [Ne ii] emission arises within
4 AU. If the dust inner rim is at ∼1 AU, as in the Ratzka et al. (2007) model, then these
models can explain the blueshift and asymmetry in the line profile since they both predict
that ∼90% of the [Ne ii] emission comes from outside 1 AU, i.e. from disk radii where the
midplane would be optically thick. A different model, also consistent with our observations,
is that from Gorti et al. (2011). From a detailed analysis of many gas emission lines from
TW Hya and using the Calvet et al. dust distribution with a static disk model, they find
that 75% of the [Ne ii] emission line comes from the surface of the optically thick dust disk
beyond the 4 AU gap.
5.2. The disk component traced by optical lines
Unlike in the case of the [Ne ii] line, the forbidden and semi-permitted optical lines
investigated here are not blueshifted with respect to the stellar radial velocity. The lack
of blueshift could have two origins:
i) the optical lines trace a wind component at disk
radii where the midplane is optically thin to the redshifted outflowing gas; OR ii) the
optical lines do not trace the photoevaporative flow but rather gas that is bound to the
star/disk system. In reference to the [OI] optical lines, this means that we do not have
an unambiguous diagnostic for the neutral wind component of the disk as we have for the
-- 29 --
ionized component. In the following, we discuss what other constraints we can place on the
disk region traced by the optical lines and what observations could discriminate between
the two scenarios presented above.
First of all, the similarity in the FWHM of the [OI] 5577, 6300, 6364 A, the [SII] at
4068 A, and the MgI] at 4571 A suggests that these lines trace similar disk radii. Based
on the centrally peaked profiles and maximum observed velocities, we argue in Sect. 4.2
that the emission could extend as close as 0.06 AU to the star, where the disk may be
magnetospherically truncated, to and possibly beyond 1 AU. The observed [OI] 6300/5577 A
line ratio of ∼7 as well as the non-detection of the [SII] doublet 6716/6730 A point to a
region of high temperature and electron density. For gas in local thermal equilibrium (LTE)
and optically thin emission we can write the [SII] 6716 and 4069 A ratio as:
L4069
L6716
=
n(2P )A4069∆E4069
n(2D)A6716∆E6716
(1)
n(2P ) and n(2D) are the upper level populations which in LTE are linked as n(2P )/n(2D) ∼
4/6 × e−14,000/T . To explain a line flux ratio higher than 6 we thus require gas hotter than
∼3,000 K. A similar argument applied to the [OI] 6300 and 5577 A lines yields a similarly
high temperature (Gorti et al. 2011). Such a high temperature can be reached within a few
AU from the star in a disk atmosphere irradiated by stellar Xray but LTE requires very high
electron densities (ne & 108cm−3) which may not be present in the photoevaporative wind
(Gorti et al. 2011). We turn now to the strong [OI] lines and on the neutral component of
the disk wind to investigate the two scenarios proposed at the beginning of this section.
In the first scenario, the [O I] emission is thermal and arises from the hot (>5,000 K)
disk surface heated by stellar Xrays. Ercolano & Owen (2010) have shown that line profiles
can be symmetric even if the gas is flowing and any blueshift in the peak emission can be
small, a few tenth of km/s, if most of the emission arises within the dust depleted inner
-- 30 --
disk6. For the specific case of TW Hya an inner dust hole of 2 AU would produce a blueshift
in the [OI] line greater than 1 km/s and slightly asymmetric profiles while a 4 AU hole would
give just a shift of 0.4 km/s and a symmetric profile pointing to the flow traced by the [OI]
emission extending well beyond 2 AU. The FWHM of the line is in both models relatively
large ∼20 km/s, a factor of 2 larger than measured. The line broadening can be reduced if
the emission in the flow is suppressed and the bound [O i] thermal component arising from
within a few AU dominates the emission. This strongly points to the second scenario in
which the [OI] lines are not tracing the neutral component of the photoevaporative wind
in the disk of TW Hya but rather the bound disk component well inside the gravitational
radius.
A variant to the second scenario is that proposed by Gorti et al. (2011). In this
model, which is specifically for TW Hya, most of the [O I] emission is not thermal but
produced from dissociation of OH molecules by stellar FUV photons leading to O atoms in
electronically excited states and their subsequent decay to the ground state. Gorti et al.
(2011) find that this mechanism gives a 6300/5577 A line ratio of ∼7, very similar to the
observed one. Because the [O i] emission comes from cooler gas closer to the disk midplane
the emitting gas would not participate to the photoevaporative flow and hence [OI] lines
would be narrower and symmetric around the stellar velocity. The [SII] line ratios are also
reproduced in this model since most of the emission arises in the hot and dense gas close to
the star within the dust depleted inner disk.
A way to directly discriminate between bound and unbound oxygen gas is
through spectroastrometry in the [OI] 6300 A line. For gas in Keplerian rotation the
6As in the case of the [Ne ii] line sub-micron dust grains are the main source of opacity.
For the [OI] 6300 A line the H column density to reach optical depth of one is as high as
2×1025 cm−2.
-- 31 --
spectroastrometric signal will be symmetric in velocity and the strongest for a slit oriented
along the disk major axis, the signal will be zero for a slit perpendicular to the major
axis (see, e.g. Fig. 12). The presence of a photoevaporative wind alters the symmetric
spectroastrometric signal and produces a shift in the flux distribution versus wavelength
even for a slit perpendicular to the disk major axis. With a combination of good
seeing (∼1′′) and high signal to-noise a few milli-arcsecond accuracy can be attained via
spectroastrometry at visible wavelengths (see e.g. Takami et al. 2003). Hence, detecting
asymmetries around 1 AU, which would produce shifts of ∼20 mas, is well within the reach
of this technique.
5.3. On determining photoevaporation rates
Our study of TW Hya represents the first attempt to identify and combine
information from multiple gas lines at infrared and optical wavelengths that could trace the
photoevaporative disk wind. Here, we summarize the observables reproduced by different
disk models as applied to TW Hya and discuss how to nail down photoevaporation rates
from protoplanetary disks.
The intensity of the [Ne ii] line at 12.81 µm can be reproduced by all models: in the
EUV-only model ionized neon atoms trace the skin of the disk while in the X-EUV and
XE-FUV models they trace a deeper, mostly neutral layer that participates to the flow.
Any asymmetry in the line profile tells us where most of the emission arises with respect to
the optically thick dust disk. A better determination of the transition from thin to thick
disk could potentially validate or invalidate the EUV/Xray photoevaporation models (see
Sect. 5.1). The strong [O i] line at 6300 A from TW Hya cannot arise from the EUV layer
because this layer is mostly ionized and produces only weak [O I] emission. While the [O I]
peak emission from several T Tauri stars in the Hartigan et al. (1995) sample presents
-- 32 --
a small blueshift pointing to a possible origin in the Xray wind, this is not the case for
TW Hya. In addition, the small FWHM of the [O I] line also argues against most of the
emission tracing the wind and rather points to bound gas close to the star. The luminosities
of the [S II] at 4068 and the Mg I] at 4571 A can be reproduced by the X-EUV primordial
disk model but with a stellar Xray luminosity that is 10 times weaker than that of TW Hya
(Table 1 from Ercolano & Owen 2010). The non-detection of the [S II] doublet at 6717 and
6731 A is also puzzling in the frame of both the EUV and the X-EUV model. For our results
to be consistent with these models, sulfur should be depleted in the disk wind. Depletion
of more refractory elements such as silicon and aluminum has been previously proposed for
the gas surrounding TW Hya (e.g., Lamzin et al. 2004). Elemental abundances therefore
add an additional uncertainty when comparing observed line fluxes to model calculations.
Alternatively the XE-FUV model developed for TW Hya explains the non-detections of the
[S II] doublet and proposes that the observed [S II] line at 4069 A arises from dense, bound
gas in the inner disk (Gorti et al. 2011). Similarly to the forbidden optical lines, the CO
line profiles and their spectroastrometric signal are consistent with bound gas in Keplerian
rotation around TW Hya (Pontoppidan et al. 2008).
More effort should be devoted in identifying gas lines that trace the predominantly
neutral winds of XE-FUV and X-EUV irradiated disks both around TW Hya as well as
around other young stars. Establishing the presence of this component would point to
flow rates 1-2 orders of magnitude larger than the 10−10 M⊙/yr predicted by EUV-only
photoevaporation models. In addition, attempts should be made to empirically measure
the ionizing flux impinging on the surface of protoplanetary disks via for instance line flux
ratios from two successive stages of ionization of a given element or from volatile elements
tracing the flow.
-- 33 --
6. Concluding remarks
In the present work we have provided observational constraints to the photoevaporative
wind from the disk of TW Hya. We confirm our previous finding that the [Ne ii] 12.81 µm
line traces the ionized component of the wind from the disk of TW Hya (Pascucci & Sterzik
2009). This would be the dominant wind component in the EUV case and a trace component
in the predominantly neutral Xray wind. From the blueshift and asymmetry in the line
profile, we estimate that most (>80%) of the [Ne ii] flux arises from disk radii where the
midplane is optically thick to the emission, meaning beyond the 1 or 4 AU dust gap inferred
from other observations. We note that depending on the observations/disk modeling this
gap may be not completely evacuated of dust (Calvet et al. 2002; Ratzka et al. 2007;
Akeson et al. 2011). We also show that the [Ne ii] emission is confined within 10 AU from
the star. Our spectroastrometry in the [Ne ii] line suggests that the wind flux is rather
spatially symmetric around the star but the current dataset is not of quality high enough
to provide additional constraints to the structure of the photoevaporative flow. We find
that optical forbidden lines such as the strong [O i] emission line at 6300 A have remarkably
different profiles from the [Ne ii] 12.81 µm line, in that they are centered at the stellar
velocity and are symmetric. The only way these lines could trace the photoevaporative
flow is if they arise from a disk region physically different from that traced by the [Ne ii]
line, specifically from within the optically thin dust gap. However, the small FWHM of the
optical lines indicates that most of the gas is likely bound, which would imply gas very
close to the star (inside the ∼1 AU critical radius for photoevaporation) if the Keplerian
velocity contributes most to the line broadening.
The above characteristics of the gaseous disk of TW Hya can be used to speculate on
the origin of its transition-like SED. One possibility for creating dust gaps in circumstellar
disks and still account for accretion onto the star is via dynamical clearing by forming giant
-- 34 --
planets (Lubow & D'Angelo 2006). For the specific case of TW Hya, partial dust clearing
is required at most out to ∼4 AU from the star (Calvet et al. 2002). Recent hydrodynamic
simulations by Zhu et al. (2011) suggest that one giant planet might account for this
moderate-size gap (see also Gorti et al. 2011) and still result in the measured accretion
rate of 10−9 M⊙/yr (Muzerolle et al. 2000; Alencar & Batalha 2002). Another possibility is
that the gap is not opened by a planet but rather by photoevaporation and we are catching
the star in the process of dispersing its disk. Recent XEUV-driven photoevaporation
models predict that there is a non-negligible amount of time (2-3×105 yr) during which
photoevaporation can carve a moderate-size gap, accretion onto the star is still detectable,
and dust grains are cleared out under the action of dust drag (Owen et al. 2011). We would
like to stress that the on-going photoevaporation from the disk of TW Hya reported here
does not exclude the existence of a giant planet and does not preclude that a forming giant
planet has dynamically cleared the inner disk. Infact, a giant planet might expedite the
onset of substantial (and detectable) photoevaporation by depleting the inner disk surface
density which reduces the stellar accretion rate and exposes the region out of the rim with
direct unattenuated flux from the star.
Besides detecting a planet at a few AU from TW Hya, there are other observables
that could help discriminating between a planet-induced and a photoevaporation-induced
disk gap. If the putative giant planet is on an eccentric orbit, one might see variations in
the dust disk structure and related dust emission on the Keplerian timescale of the planet,
a few years. Regardless of the planet's eccentricity, dynamical clearing would result in a
sharp transition in surface density at the outer gap edge which could be imaged with the
Atacama Large Millimeter Array.
IP is pleased to acknowledge support from the National Science Foundation (NSF)
through an Astronomy & Astrophysics research grant (AST0908479).
IP thanks E.
-- 35 --
Flaccomio for making available the VISIR spectra of Titan and K. Pontoppidan for valuable
discussions. RDA acknowledges support from the Science & Technology Facilities Council
(STFC) through an Advanced Fellowship (ST/G00711X/1). SHPA acknowledges support
from Fapemig and CAPES, brazilian research agencies. This research used the ALICE High
Performance Computing Facility at the University of Leicester. Some resources on ALICE
form part of the DiRAC Facility jointly funded by STFC and the Large Facilities Capital
Fund of BIS.
Facilities: VLT (VISIR), 1.5m ESO (FEROS).
-- 36 --
Slit center+/-(FWHM/2)
Slit-Source center
63 90
Sky Position Angle (degree)
153 180
243 270
333
)
l
e
x
p
(
i
n
o
i
t
c
e
r
i
d
d
e
s
r
e
p
s
d
i
s
s
o
r
c
e
c
r
u
o
S
-
t
i
l
S
4
2
0
-2
-4
0
Fig. 1. -- Difference in pixels between the location of TW Hya and the center of the slit
before each spectroscopic integration. The x-axis gives the position angle in degrees. Vertical
errorbars centered at y=0 show the extension of the slit image (FWHM) in cross-dispersion
direction. Except for PA=63◦ TW Hya was well positioned inside the slit.
-- 37 --
PA = 0o
PA = 180o
PA = 270o
PA = 90o
PA = 333o
]
y
J
[
x
u
F
l
6
4
2
0
-60
-40
-20
0
0
Velocity [km/s]
20
40
60
Fig. 2. -- The five flux calibrated spectra of TW Hya around the [Ne ii] emission line. The
x-axis gives the velocity in the stellocentric frame.
-- 38 --
VISIR spectrum of Titan
12.80
12.81
Wavelength [µm]
12.82
12.83
40
30
20
10
0
4
2
0
-2
y
t
i
s
n
e
t
n
I
]
s
/
m
k
[
v
∆
-4
12.79
Fig. 3. -- VISIR spectra of Titan demonstrating the accuracy of the dispersion solution. The
upper panel marks in (green) dashed-lines the C2H2 transitions from the HITRAN database.
This spectrum of Titan (we analyzed two spectra) has been shifted to match the single C2H2
line at ∼12.803µm. The lower panel gives the velocity difference in km/s for lines marked in
the upper panel. The location of the [Ne ii] line is also shown (black vertical dashed line).
The two single C2H2 lines are marked with arrows.
'X' symbols give shifts when the rest
frame is computed from the single line at ∼12.819µm, while diamonds from the single line
at ∼12.803µm. The standard deviation of the velocity difference, which gives the relative
uncertainty, is 1 km/s. The absolute uncertainty is a factor of two larger (see text).
-- 39 --
153
180
333
270
0
63
180
270
333
1.0
1.1
0
1.2
Airmass
1.3
90
243
90
1.4
50
45
40
35
30
25
-4
-5
-6
-7
W
E
k
a
e
p
V
x
u
F
l
7
6
5
4
3
2
0.8
0.6
0.4
0.2
m
u
u
n
i
t
n
o
C
Fig.
4. -- [Ne ii]
line EW (A), peak centroid (km/s in the stellocentric frame), flux
(×10−14erg cm−2 s−1), and adjacent continuum (Jy) versus the airmass at which the spec-
troscopic observations were carried out. PAs are also overplotted. Note the anti-correlation
between the [Ne ii] flux and adjacent continuum with airmass suggesting that most of the
observed variations in these quantities are related to flux calibration uncertainties rather
than to intrinsic source variability. Also note that for PA=63◦, which shows the lowest EW
and the largest blueshift, the source was not well centered in the slit (see Fig. 1).
-- 40 --
30
25
20
15
10
5
0
-6
-4
vstar(CaI6439)
vstar(OI6364)
4
6
-2
0
Velocity OI6300 - vstar [km/s]
2
Fig. 5. -- Histogram of the velocity shifts of the [O i] 6300 A line with respect to the stellar
velocity measured via the photospheric CaI 6439 (black dashed line) and the [O i] 6364 A
(red solid line) lines. The CaI 6439 A and the [O i] 6364 A fall on the same echelle order
while the [O i] 6300 A line falls on a different echelle order. The fact that the two histograms
look very similar and have the same median and standard deviation demonstrate that the
[O i] 6300 A line is centered at the stellar velocity and the small blueshift between the [O i]
6300 and the CaI 6439 A line is just due to small differences in the dispersion solutions of
the two echelle orders.
-- 41 --
PA=90o
Original
Mirror
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
0.1
0.0
-0.1
k
a
e
p
o
t
m
r
o
N
r
o
r
r
i
M
-
l
a
n
g
i
r
i
O
-100
-50
0
Velocity [km/s]
50
100
Fig. 6. -- [Ne ii] line profiles from TW Hya. The upper panel gives the original (black)
and mirror-shifted (red) profile for PA=90◦. Note that the continuum is subtracted and the
emission is normalized to the peak. The lower panel shows the difference curves (original-
mirror) for all eight PAs. The color coding is as follows:
light black for PA 180◦, red for
90◦, green for 0◦, blue for 270◦, yellow for 153◦, cyan for 63◦, violet for 243◦, and orange for
333◦. The average difference curve is overplotted using a thick black line. Note that most
difference curves (as well as the average difference curve) have an excess on the blue side,
peaking at about -20 km/s. This shows that the line is asymmetric with a blue wing due to
photoevaporating gas accelerating toward the observer.
-- 42 --
Lines from Titan
Original
Mirror
a
m
g
s
3
i
a
m
g
s
3
i
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
0.1
0.0
-0.1
k
a
e
p
o
t
m
r
o
N
r
o
r
r
i
M
-
l
a
n
g
i
r
i
O
-40
-20
0
Velocity [km/s]
20
40
Fig. 7. -- Line profiles from Titan (PID 083.C-0883). The continuum near the line is removed
and lines are normalized to the peak emission as for TW Hya (also note the same y scale as
in Fig. 6). Lower panel: difference curves (original-mirror profiles) for 3 emission lines from
two different spectra of Titan. This figure shows that the asymmetry seen in the [Ne ii] line
profiles of TW Hya is not an instrumental artifact.
-- 43 --
Original
Mirror
1.0
0.8
0.6
0.4
0.2
0.0
-0.2
0.1
0.0
-0.1
k
a
e
p
o
t
m
r
o
N
r
o
r
r
i
M
-
l
a
n
g
i
r
i
O
-100
-50
0
Velocity [km/s]
50
100
Fig. 8. -- Asymmetry analysis for the [O i] 6300A emission lines of TW Hya. The upper
panel gives the original (black) and mirror-shifted (red) profile from one of the optical spectra
obtained by Alencar & Batalha (2002). The lower panel shows the difference curves (original-
mirror) from the 10 spectra obtained in February 2000 (Alencar & Batalha 2002), different
colors are for spectra obtained in different days. Note that unlike the [Ne ii] line at 12.81µm
the [O i] line at 6300A is not blueshifted and has a symmetric profile.
-- 44 --
k
a
e
p
o
t
m
r
o
N
0.10
0.05
0.00
-0.05
-0.10
EUV i=3o
Xrays i=3o
EUV i=10o
Xrays i=10o
-60
-40
-20
0
Velocity [km/s]
20
40
60
Fig. 9. -- Comparison of difference curves (original-mirror profiles) for predicted and ob-
served [Ne ii] line profiles. Black solid and dashed lines are the difference curves from the
EUV photoevaporation model, while red solid and dashed lines are from the Xray photoe-
vaporation model for disks inclined by 10 and 3◦ respectively. All models assume that the
disk midplane is optically thick. The blue line is the median of the difference curves from
the observed [Ne ii] profiles of TW Hya (see also Fig. 6) with the standard deviation of the
curves overplotted.
-- 45 --
Parallel orientation: 0,270,333,63
1.0
0.5
0.0
-0.5
-1.0
Anti-Parallel orientation: 180,90,153,243
1.0
0.5
0.0
-0.5
-1.0
l
s
e
x
P
i
l
s
e
x
P
i
-20
-10
Velocity [km/s]
0
10
-20
-10
Velocity [km/s]
0
10
Fig. 10. -- Centroid in pixels versus velocities in the stellocentric frame for the parallel (left
panel) and anti-parallel (right panel) slit orientations. The color coding is as follows: black
for 0 and 180◦, red for 270 and 90◦, green for 333 and 153◦, and blue for 63 and 243◦.
-- 46 --
Simulated [NeII] emission from the disk of TW Hya
100
50
0
0
-50
100
50
0
0
-50
v = -10 km/s
v = +1 km/s
-100
-100
-50
0
0
distance [mas]
50
-100
100
-100
-50
0
0
distance [mas]
50
100
]
s
a
m
[
e
c
n
a
t
s
d
i
Fig. 11. -- Modelled [Ne ii] emission for two velocity channels, -10 km/s and +1 km/s. The
asymmetry is mainly due to the combination of disk inclination (4◦) and Keplerian rotation.
The wind is mainly symmetric but introduces deviations from the spectroastrometric Ke-
plerian profiles, see text. The TW Hya disk position angle is 333◦. On the left panel the
black rim N-W is a factor of ∼3 brighter than the grey emission S-E. On the right panel the
contrast between the bright emission (now S-E) and the weaker emission (N-W) is about 5
but occurs at a much smaller spatial scale. These images are convolved with the VISIR PSF
when simulating the spectro-astrometric signal.
-- 47 --
0o
270o
333o
63o
Keplerian+wind
Only wind
0.2
0.1
0.0
l
]
s
e
x
P
i
[
f
f
i
D
-0.1
-0.2
-0.3
-20
-10
0
Velocity [km/s]
10
20
Fig. 12. -- Spectroastrometric signal (parallel minus antiparallel centroids) predicted by a
disk photoevaporated by stellar EUV photons (solid lines). Dashed lines show only the wind
component, i.e. when Keplerian rotation is artificially set to zero. The plot shows that
the largest spectroastrometric signal from the disk wind is expected for PA=63◦, the slit
orientation perpendicular to the disk major axis. One pixel on VISIR is 127 mas.
-- 48 --
0o
270o
333o
63o
1.0
0.5
0.0
-0.5
-1.0
1.0
0.5
0.0
-0.5
-1.0
l
]
s
e
x
P
i
[
f
f
i
D
l
]
s
e
x
P
i
[
f
f
i
D
-20
-10
Velocity [km/s]
0
10
Fig. 13. -- Spectroastrometric signal observed (diamonds) and predicted (lines, see Fig. 12
for color coding). The y-axis gives the difference in pixels between a position angle and its
antiparallel position angle (e.g., 0◦corresponds to 0◦-180◦). One pixel on VISIR is 127 mas,
therefore the level of asymmetry we detect has a radial extension of ∼1.3 AU.
-- 49 --
Table 1. Summary of the TW Hya observations. The observing time (in U.T.), the
airmass, and the heliocentric radial velocity corrections (vhelio) are given at the beginning
and at the end of each spectroscopic exposure. On-source exposure times for TW Hya were
set to 1 h, for standard stars exposures varied between 240 and 360 s. In the second night,
we observed only one standard star for PA=333◦.
PAa
(◦)
0
180
270
90
333
153
63
243
Date
U.T.
Airmass
yyyy-mm-dd
hh:mm
vhelio
(km/s)
Calibrator
Airmass
2010-02-23
02:12/03:41
1.3/1.1
11.53/11.40
HD 90957
2010-02-23
03:54/05:23
1.1/1.0
11.38/11.23
HD 90957
2010-02-23
06:27/07:54
1.0/1.2
11.11/10.97 HD 101666
2010-02-23
08:25/09:42
1.3/1.5
10.93/10.84 HD 101666
2010-02-24
02:45/04:15
1.2/1.1
11.11/10.97
HD 90957
2010-02-24
04:28/05:58
1.0/1.0
10.95/10.79
2010-02-24
06:24/07:51
1.0/1.2
10.74/10.70
2010-02-24
08:07/09:38
1.2/1.7
10.58/10.47
--
--
--
1.3
1.0
1.0
1.5
1.2
--
--
--
aThe position angle (PA) is measured from N toward E.
Table 2. Main properties of the TW Hya star/disk system inferred from other studies and
utilized in this paper.
Property
Value
Reference
Stellar mass
Distance
0.7 M⊙
51±4 pc
1
2
Disk inclination
4.3±1◦,7±1◦,5-6◦
3,4,5
Disk PA
332±10◦, 335±2◦
Stellar ionizing flux
1041 phot/s
3,5
6
References. -- (1) Muzerolle et al. 2000; (2) Mamajek
2005; (3) Pontoppidan et al. 2008; (4) Qi et al. 2004; (5)
Hughes et al. 2011; (6) Pascucci & Sterzik 2009
-- 50 --
Table 3. Main parameters of the [Ne ii] line and adjacent continuum for different position
angles (PA).
PA
(◦)
0
180
270
90
333
153
63
243
EW
(A)
FWHM
vpeak
Line Flux
Continuum
(km/s)
(km/s)
(×10−14erg cm−2 s−1)
(Jy)
43±1
15.2±0.4
-5.0±0.2
47.1±0.8
16.9±0.4
-4.9±0.1
44±1
36±1
47±1
15.0±0.4
-4.8±0.2
15.8±0.5
-5.9±0.2
15.9±0.4
-6.2±0.2
47.7±0.8
16.6±0.3
-3.9±0.3
30±1
35±1
17.5±0.7
-7.5±0.3
17.2±0.6
-3.9±0.3
2.6±0.1
6.5±0.2
4.4±0.1
3.0±0.1
3.2±0.1
--
--
--
0.3
0.8
0.6
0.5
0.4
--
--
--
Note. -- Errors on the EWs are computed using a Monte Carlo approach: we
added a normally distributed noise to the spectrum (following the error at each
wavelength) and give as error the standard deviation of the EW distribution of 1000
such spectra. The errors on the FWHM, peak emission, and line flux are computed
from the gaussian fit to the [Ne ii] line taking into account the uncertainty at each
wavelength. Note that the absolute radial velocity accuracy reachable with VISIR
is ∼2 km/s (Sect. 3.1), larger than the uncertainties on the [Ne ii] peak velocities
derived from the gaussian fit.
-- 51 --
Table 4. Line EWs and upper limits for other optical lines predicted to trace the
photoevaporative wind.
Species Wavelength
(A)
Terms
a
Aul
(s−1)
OI
OI
OI
OII
MgI
NII
SII
SII
SII
SII
5577.3387
1.26(0)
1D-1S
6300.304
5.65(-3)
3P-1D
6363.777
1.82(-3)
3P-1D
7330.73
5.34(-2)
2D-2P
<0.04
4571.0956
2.54(2)
1S-3P
0.08
6583.460
2.92(-3)
3P-1D <0.007
4068.600
3.41(-1)
4076.349
1.34(-1)
4S-2P
4S-2P
0.09
<0.03
6716.440
4.66(-4)
4S-2D
<0.01
6730.810
4.26(-4)
4S-2D <0.008
EW
(A)
0.09
0.47
0.15
veiling
Luminositya
(L⊙)
1.5(-6)
1.1(-5)
3.2(-6)
<9.1(-7)
6.4(-7)
<1.5(-7)
1.4(-6)
<4.8(-7)
<2.3(-7)
<1.9(-7)
0.5
0.6
0.4
0.5
0.5
0.4
2
2
0.5
0.6
aThe transition probability from NIST and the line luminosity are given as a(b)
meaning a×10b.
Note. -- Wavelengths are in air. EWs, veilings, and corresponding line lumi-
nosities are median values. Uncertainties on the measured EWs are about ∼10%.
Because of night-to-night variations in the source brightness and in the veiling the
luminosities reported here have an uncertainty of ∼50% (see Sect. 4.2).
-- 52 --
REFERENCES
Adams, F. C., Hollenbach, D., Laughlin, G., & Gorti, U. 2004, ApJ, 611, 360
Akeson, R. L., et al. 2011, ApJ, 728, 96
Alencar, S. H. P., & Batalha, C. 2002, ApJ, 571, 378
Alexander, R. D. 2008, MNRAS, 391, L64
Alexander, R. D., Clarke, C. J., & Pringle, J. E. 2006a, MNRAS, 369, 216
-- . 2006b, MNRAS, 369, 229
Armitage, P. J. 2010, ArXiv e-prints
Bailey, J. 1998, MNRAS, 301, 161
Calvet, N., D'Alessio, P., Hartmann, L., Wilner, D., Walsh, A., & Sitko, M. 2002, ApJ, 568,
1008
Calvet, N., Hartmann, L., & Strom, S. E. 2000, Protostars and Planets IV, 377
Cieza, L. A., Swift, J. J., Mathews, G. S., & Williams, J. P. 2008, ApJ, 686, L115
Clarke, C. J., Gendrin, A., & Sotomayor, M. 2001, MNRAS, 328, 485
Currie, T., Lada, C. J., Plavchan, P., Robitaille, T. P., Irwin, J., & Kenyon, S. J. 2009,
ApJ, 698, 1
Dullemond, C. P., Hollenbach, D., Kamp, I., & D'Alessio, P. 2007, Protostars and Planets
V, 555
Ercolano, B., & Clarke, C. J. 2010, MNRAS, 402, 2735
Ercolano, B., Drake, J. J., Raymond, J. C., & Clarke, C. C. 2008, ApJ, 688, 398
-- 53 --
Ercolano, B., & Owen, J. E. 2010, MNRAS, 406, 1553
Font, A. S., McCarthy, I. G., Johnstone, D., & Ballantyne, D. R. 2004, ApJ, 607, 890
Glassgold, A. E., Najita, J. R., & Igea, J. 2007, ApJ, 656, 515
Gorti, U., Dullemond, C. P., & Hollenbach, D. 2009, ApJ, 705, 1237
Gorti, U., & Hollenbach, D. 2009, ApJ, 690, 1539
Gray, D. F. 1980, ApJ, 235, 508
Hartigan, P., Edwards, S., & Ghandour, L. 1995, ApJ, 452, 736
Hartmann, L., Calvet, N., Gullbring, E., & D'Alessio, P. 1998, ApJ, 495, 385
Hayashi, C., Nakazawa, K., & Nakagawa, Y. 1985, in Protostars and Planets II, ed.
D. C. Black & M. S. Matthews, 1100 -- 1153
Herczeg, G. J., Najita, J. R., Hillenbrand, L. A., & Pascucci, I. 2007, ApJ, 670, 509
Hollenbach, D., & Gorti, U. 2009, ApJ, 703, 1203
Hollenbach, D., Johnstone, D., Lizano, S., & Shu, F. 1994, ApJ, 428, 654
Hollenbach, D., & McKee, C. F. 1989, ApJ, 342, 306
Hollenbach, D. J., Yorke, H. W., & Johnstone, D. 2000, Protostars and Planets IV, 401
Horne, K. 1986, PASP, 98, 609
Hughes, A. M., Wilner, D. J., Andrews, S. M., Qi, C., & Hogerheijde, M. R. 2011, ApJ,
727, 85
Kaufl, H. U. 2006, in Planetary Nebulae Beyond the Milky Way, ed. L. Stanghellini,
J. R. Walsh, & N. G. Douglas, 201 -- +
-- 54 --
Kim, K. H., et al. 2009, ApJ, 700, 1017
Lagage, P. O., et al. 2004, The Messenger, 117, 12
Liffman, K. 2003, PASA, 20, 337
Lubow, S. H., & D'Angelo, G. 2006, ApJ, 641, 526
Mamajek, E. E. 2005, ApJ, 634, 1385
Muzerolle, J., Calvet, N., Briceno, C., Hartmann, L., & Hillenbrand, L. 2000, ApJ, 535, L47
Muzerolle, J., Hartmann, L., & Calvet, N. 1998, AJ, 116, 455
Najita, J. R., Carr, J. S., Strom, S. E., Watson, D. M., Pascucci, I., Hollenbach, D., Gorti,
U., & Keller, L. 2010, ApJ, 712, 274
Natta, A., Testi, L., Calvet, N., Henning, T., Waters, R., & Wilner, D. 2007, Protostars
and Planets V, 767
Owen, J. E., Ercolano, B., & Clarke, C. J. 2011, MNRAS, 412, 13
Owen, J. E., Ercolano, B., Clarke, C. J., & Alexander, R. D. 2010, MNRAS, 401, 1415
Pascucci, I., & Sterzik, M. 2009, ApJ, 702, 724
Pascucci, I., & Tachibana, S. 2010, The Clearing of Protoplanetary Disks and of the
Protosolar Nebula (Apai, D. A. & Lauretta, D. S.), 263 -- 298
Pontoppidan, K. M., Blake, G. A., van Dishoeck, E. F., Smette, A., Ireland, M. J., &
Brown, J. 2008, ApJ, 684, 1323
Qi, C., et al. 2004, ApJ, 616, L11
-- 55 --
Ratzka, T., Leinert, C., Henning, T., Bouwman, J., Dullemond, C. P., & Jaffe, W. 2007,
A&A, 471, 173
Rothman, L. S., & Gordon, I. E. 2009, in 64th International Symposium On Molecular
Spectroscopy
Rucinski, S. M., et al. 2008, MNRAS, 391, 1913
Sicilia-Aguilar, A., Henning, T., & Hartmann, L. W. 2010, ApJ, 710, 597
Weintraub, D. A., Kastner, J. H., & Bary, J. S. 2000, ApJ, 541, 767
Whelan, E., & Garcia, P. 2008, in Lecture Notes in Physics, Berlin Springer Verlag, Vol.
742, Jets from Young Stars II, ed. F. Bacciotti, L. Testi, & E. Whelan, 123 -- +
Yasui, C., Kobayashi, N., Tokunaga, A. T., Saito, M., & Tokoku, C. 2009, ApJ, 705, 54
Zhu, Z., Nelson, R. P., Hartmann, L., Espaillat, C., & Calvet, N. 2011, ApJ, 729, 47
This manuscript was prepared with the AAS LATEX macros v5.2.
|
1211.5148 | 1 | 1211 | 2012-11-21T21:04:18 | Millimeter Emission Structure in the first ALMA Image of the AU Mic Debris Disk | [
"astro-ph.EP",
"astro-ph.SR"
] | We present 1.3 millimeter ALMA Cycle 0 observations of the edge-on debris disk around the nearby, ~10 Myr-old, M-type star AU Mic. These observations obtain 0.6 arcsec (6 AU) resolution and reveal two distinct emission components: (1) the previously known dust belt that extends to a radius of 40 AU, and (2) a newly recognized central peak that remains unresolved. The cold dust belt of mass about 1 lunar mass is resolved in the radial direction with a rising emission profile that peaks sharply at the location of the outer edge of the "birth ring" of planetesimals hypothesized to explain the midplane scattered light gradients. No significant asymmetries are discerned in the structure or position of this dust belt. The central peak identified in the ALMA image is ~6 times brighter than the stellar photosphere, which indicates an additional emission process in the inner regions of the system. Emission from a stellar corona or activity may contribute, but the observations show no signs of temporal variations characteristic of radio-wave flares. We suggest that this central component may be dominated by dust emission from an inner planetesimal belt of mass about 0.01 lunar mass, consistent with a lack of emission shortward of 25 microns and a location <3 AU from the star. Future millimeter observations can test this assertion, as an inner dust belt should be readily separated from the central star at higher angular resolution. | astro-ph.EP | astro-ph |
accepted by ApJ Letters: November 17, 2012
Millimeter Emission Structure in the first ALMA Image of the
AU Mic Debris Disk
Meredith A. MacGregor1, David J. Wilner1, Katherine A. Rosenfeld1, Sean M. Andrews1,
Brenda Matthews2, A. Meredith Hughes3, Mark Booth2,4, Eugene Chiang3,
James R. Graham3,5, Paul Kalas3,6, Grant Kennedy7, Bruce Sibthorpe8
ABSTRACT
We present 1.3 millimeter ALMA Cycle 0 observations of the edge-on debris
disk around the nearby, ∼10 Myr-old, M-type star AU Mic. These observations
obtain 0.′′6 (6 AU) resolution and reveal two distinct emission components: (1) the
previously known dust belt that extends to a radius of 40 AU, and (2) a newly
recognized central peak that remains unresolved. The cold dust belt of mass
∼ 1 MMoon is resolved in the radial direction with a rising emission profile that
peaks sharply at the location of the outer edge of the "birth ring" of planetesimals
hypothesized to explain the midplane scattered light gradients. No significant
asymmetries are discerned in the structure or position of this dust belt. The
central peak identified in the ALMA image is ∼ 6 times brighter than the stellar
photosphere, which indicates an additional emission process in the inner regions
of the system. Emission from a stellar corona or activity may contribute, but
the observations show no signs of temporal variations characteristic of radio-
wave flares. We suggest that this central component may be dominated by dust
1Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
2Herzberg Institute of Astrophysics, 5072 West Saanich Road, Victoria, BC V9E 2E7, Canada
3Department of Astronomy, 601 Campbell Hall, University of California, Berkeley, CA 94720, USA
4Deptarment of Physics & Astronomy, University of Victoria, 3800 Finnerty Rd., Victoria, BC, V8P 5C2,
Canada
5Dunlap Institute for Astronomy & Astrophysics, University of Toronto, Toronto, ON, Canada
6SETI Institute, 189 Bernardo Ave., Mountain View, CA 94043
7Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA, UK
8SRON Netherlands Institute for Space Research, NL-9747 AD Groningen, The Netherlands
-- 2 --
emission from an inner planetesimal belt of mass ∼ 0.01 MMoon, consistent with a
lack of emission shortward of 25 µm and a location .3 AU from the star. Future
millimeter observations can test this assertion, as an inner dust belt should be
readily separated from the central star at higher angular resolution.
Subject headings:
dividual (AU Microscopii) -- submillimeter: planetary systems
circumstellar matter -- planet-disk interactions -- stars: in-
1.
Introduction
Debris disks are created by the collisional erosion of planetesimals, the building blocks
of planetary systems. These collisions continuously generate dust grains with a range of
sizes that are detected with astronomical measurements from optical to radio wavelengths.
Resolved observations of nearby debris disks are instrumental in advancing our understanding
of these systems. At a distance of 9.91 ± 0.10 pc (van Leeuwen 2007), the M1 star AU Mic
hosts one of the closest and best studied debris disks. The detection of submillimeter emission
(Liu et al. 2004) from this ∼10 Myr-old system in the β Pic moving group (Zuckerman et al.
2001) was followed quickly by the discovery of an edge-on disk seen in scattered starlight
(Kalas et al. 2004). Subsequent work has characterized the scattered light in great detail,
exploiting its proximity to constrain its radial and vertical structure (Liu 2004; Krist et al.
2005; Metchev et al. 2005; Graham et al. 2007; Fitzgerald et al. 2007).
Observations of dust emission at (sub)millimeter wavelengths provide important, com-
plementary information about debris disk structures. Unlike the small grains probed at op-
tical and near-infrared wavelengths that react strongly to stellar radiation and wind forces,
the large grains that dominate the millimeter-wave emission have dynamics more like the
parent planetesimals. As a result, long-wavelength images trace best the location and dis-
tribution of the larger colliding bodies (Wyatt 2006), and potentially also the signatures
of planets that interact with them (Ertel et al. 2012). These size-dependent dust dynamics
manifest beautifully in the edge-on AU Mic disk. Resolved millimeter-wave observations
show an emission belt within the extended optical disk that peaks near a radius of 35 AU,
where the midplane scattered light profile steepens dramatically (Wilner et al. 2012). These
features are elegantly explained by the presence of a "birth ring" of planetesimals at that
location, where small grains released in a collisional cascade are launched into an extended
halo (Strubbe & Chiang 2006; Augereau & Beust 2006).
With the advent of the Atacama Large Millimeter Array (ALMA), the millimeter emis-
sion in nearby debris disks can be imaged in much greater detail (e.g., Boley et al. 2012). In
-- 3 --
this Letter, we present new, sub-arcsecond resolution ALMA Cycle 0 observations of AU Mic
at λ = 1.3 mm. The ALMA data provide substantially improved constraints on the locations
of colliding planetesimals in the AU Mic disk and help shed light on the processes that may
be shaping the planetesimal distribution. They also reveal a previously unknown, centrally
located emission feature.
2. Observations
AU Mic was observed by ALMA with its Band 6 receivers over four 2 hour-long "schedul-
ing blocks" (SBs) in 2012 Apr and Jun. Table 1 summarizes the observations. The 16-20
operational 12-m antennas were arranged to span baseline lengths of 21 -- 402 m (correspond-
ing to a maximum resolution of ∼0.′′6). The correlator was configured to optimize continuum
sensitivity, processing two polarizations in four 2 GHz-wide basebands, each with 128 spectral
channels, centered at 226, 228, 242, and 244 GHz. In each SB, we interleaved observations
of AU Mic (pointing center α = 20h45m09.s34, δ = −31◦20′24.′′09, J2000, within 1′′ of the star
position at all epochs) with the nearby quasar J2101−295.
The data from each SB were calibrated independently within the CASA software package.
After applying system temperature measurements and phase corrections from the water
vapor radiometers, the data were flagged and averaged into 6.048 s integrations. A calibration
of the spectral response of the system was determined from observations of J1924−292, and
complex gain variations induced by atmospheric and instrumental effects were corrected using
observations of J2101−295. The absolute flux calibration was derived from observations of
Neptune: a mean calibration was applied to all basebands, with a systematic uncertainty
of ∼10% (see §3.3). To generate an image at the mean frequency, 235 GHz (1.28 mm), we
Fourier inverted the calibrated visibilities with natural weighting and performed a multi-
frequency synthesis deconvolution with the CLEAN algorithm. The visibilities were further
reduced by spectrally averaging over the central 112 channels in each baseband and re-
weighted by the observed scatter.
3. Results and Analysis
3.1.
Image of 1.3 mm Dust Continuum Emission
Figure 1 shows an image of the λ = 1.3 mm emission from SB-4 (with the most antennas
and best weather conditions), with synthesized beam 0.′′80 × 0.′′69 (8 × 7 AU), p.a. 49◦, and
rms of 30 µJy beam−1. An image constructed from all 4 SBs is consistent but noisier, which
-- 4 --
Fig. 1. -- ALMA image of the 1.3 mm continuum emission from AU Mic. The ellipse in the
lower left corner represents the 0.′′80 × 0.′′69 (8 × 7 AU) synthesized beam.
we attribute to systematic calibration issues resulting from the poorer weather conditions of
the earlier observations. The emission is confined to a narrow band with aspect ratio >10:1,
with an orientation consistent with the scattered light disk. The emission is not resolved
in the direction perpendicular to the elongation. There are clear peaks near both extrema
and in the middle of the structure (detected at all four epochs). The emission is marginally
brighter at the northwest end than the southeast end, and shows small undulations along its
length, though none of these variations are significant. We interpret the observed structure
as a superposition of two components: (1) the nearly edge-on dust belt with limb-brightened
ansae, and (2) a new, distinct, and compact feature located at the center of the belt.
3.2. Modeling Formalism
Building on the phenomenological methodology of Wilner et al. (2011, 2012) to analyze
resolved millimeter emission from debris disks, we construct a parametric model to quantify
the observed properties of the AU Mic emission. We consider two model components: a
vertically thin, axisymmetric "outer" belt, and an additional source to account for the central
-- 5 --
peak. The belt component is informed by models of the scattered light that show the disk
midplane within 50 AU is remarkably straight, .0.◦5 from edge-on, and thin (FWHM ∼0.′′3).
We assume the belt is viewed at an inclination of 89.◦5. The belt is characterized as an annulus
with (unprojected) radial intensity Iν(r) ∝ rx for ri < r < ro, with a normalization defined
by Fbelt = R Iν dΩ, a center determined by offsets (relative to the pointing center) {∆α, ∆δ},
and an orientation described by a position angle (PA). We treat the central component as
a circular Gaussian with mean ∆rcen, variance σ2
cen (half width at half maximum Rcen =
√2 ln 2 σcen), and flux density Fcen. The mean ∆rcen is defined as a radial shift from the belt
center in the plane of the belt. We also include power-law spectral scalings between the 4
basebands for each component, denoted αbelt and αcen, where Fν ∝ να.
For a given parameter set, we compute four synthetic visibility sets sampled at the same
spatial frequencies observed by ALMA, corresponding to the spectrally averaged basebands
(at 226, 228, 242, and 244 GHz). By fitting the visibility data directly, we are not sensitive
to the non-linear effects of deconvolution, and take advantage of the full range of available
spatial frequencies. The fit quality is quantified by a likelihood metric, L, determined from
the χ2 values summed over the real and imaginary components at all spatial frequencies
(lnL = −χ2/2). A Monte Carlo Markov Chain (MCMC) approach was utilized to char-
acterize the multi-dimensional parameter space of this model and determine the posterior
probability distribution functions for each parameter. We used the affine-invariant ensemble
sampler proposed by Goodman & Weare (2010), in a locally-modified version of the paral-
lelized implementation described by Foreman-Mackey et al. (2012), to compute likelihood
values for ∼106 MCMC trials. Uniform priors were assumed for all parameters, with bounds
imposed to ensure that the model was well-defined: {Fbelt, Fcen, σ2
cen} ≥ 0, and 0 ≤ ri < ro.
3.3. Results of Model Fits
The best-fit parameter values and their 68% uncertainties determined from the marginal-
ized posterior probability distributions are listed in Table 2. The data and best-fit model
are compared in the image plane in Figure 2; there are no significant residuals. The best-fit
model has a reduced χ2 = 1.37 (905,920 independent datapoints, 12 free parameters). The
modeling procedure was performed on each SB individually and the full dataset (all 4 SBs
together). The results were entirely consistent, although the parameter uncertainties were
notably smaller from the superior SB-4 dataset alone, and we focus on those results.
Most parameters are determined with high precision. We find good agreement of the
outer belt parameters {Fbelt, ri, ro} with the less well-constrained fits of Wilner et al. (2012),
and on the disk PA from measurements of scattered starlight (e.g., Krist et al. 2005). We
-- 6 --
Fig. 2. -- (left) The observed 1.3 mm emission from AU Mic, (center) the best-fit model
(see §3.3), and (right) the imaged residuals. Contours are drawn at 4 σ (120 µJy beam−1)
intervals.
measure a flat spectrum for the outer belt (αbelt ≈ 0) across the 4 basebands, which corre-
sponds to the difference between the spectral slopes of AU Mic and Neptune (αNeptune ≈ 2.1),
consistent with data from 350 µm to 1.3 mm (Wilner et al. 2012).
The central emission peak is detected with high confidence at Fcen = 320 µJy (>10 σ
brighter than the outer belt at that location). It is unresolved, with Rcen ≤ 3.0 AU (3 σ),
and positionally coincident with the outer belt center: ∆rcen ≤ 1.9 AU (3 σ). Regarding the
outer belt, the most notable result is that the models strongly favor rising emission profiles
with large, positive gradients: x ≈ 2.3 ± 0.3. Models with the standard assumption of x < 0
produce significant residuals, under-predicting the intensities at ±1-2′′ from the belt center.
Because of the steep increase in the emission profile, there is only a weak constraint on the
inner edge of the outer belt. The best-fit ri deviates from 0 at the ∼2 σ level: the 3 σ limit
is ri ≤ 21 AU.
4. Discussion
We have presented new, sub-arcsecond resolution ALMA observations of 1.3 mm emis-
sion from the AU Mic debris disk and analyzed the data with a simple parametric model.
This emission is resolved into two distinct components: (1) an edge-on outer belt with an
emission profile that rises with radius out to 40 AU, and (2) an unresolved peak at the
center of the outer belt. This distribution is more complex than the single, narrow ring often
-- 7 --
assumed for debris disks. However, it has some similarities to other nearby resolved systems,
like ǫ Eri (Backman et al. 2009) or HR 8799 (Su et al. 2009), that show an inner component
inferred from excess infrared emission, separate from an extended and colder outer belt.
4.1. The Central Emission Peak
The stellar photosphere is much fainter than the central peak noted in Figure 1. A
NextGen stellar model (Hauschildt et al. 1999) with Teff = 3720 K, L∗ = 0.11 L⊙, and
M∗ = 0.6 M⊙ (e.g., Metchev et al. 2005; Chen et al. 2005) that matches the AU Mic pho-
tometry from 0.4 -- 25 µm contributes only F∗ = 52 µJy at 1.3 mm, ∼6× fainter than observed.
However, AU Mic is an active star that exhibits radio-wave bursts. In quiescence, observa-
tions find <120 µJy at 3.6 cm (White et al. 1994), and the contribution at 1.3 mm from hot
coronal plasma seen in X-rays is unlikely to be significant (though better spectral constraints
are desirable, see Leto et al. 2000). Flares are detected from AU Mic at ∼200-1200 µJy at
6 cm (Bower et al. 2009), but this non-thermal emission is much weaker at 1.3 mm. While
the unknown variability makes any extrapolation to 1.3 mm problematic, the temporal prop-
erties of the ALMA emission provide additional information. Radio-wave flares have fast
decay times, of order an hour (Kundu et al. 1987); but, the mm-wave peak persists at a
consistent intensity in all four ALMA observations, within uncertainties that are typically
2 − 3× larger than for SB-4, spanning timescales from 1 hour (within SB-4) to 2 months
(SB-1 to SB-4). Unfortunately, the spectral index (αcen) constraints are not good enough to
be diagnostic. We suspect that stellar emission is too weak and too ephemeral to be respon-
sible for the 1.3 mm peak, but the available data does not allow for a firm determination of
its contribution.
Alternatively, the central emission peak could be produced by dust in a distinct (un-
resolved) planetesimal belt located close to the star. In §3.3, we constrained the extent of
this peak to Rcen ≤ 3 AU (3 σ), inside the inner working angle (0.′′8 ≈ 8 AU) of all previous
high resolution imaging of scattered light (Krist et al. 2005; Fitzgerald et al. 2007). Rough
models of the spectral energy distribution (SED) from the ALMA central peak can help
assess the feasibility that it originates in an inner dust belt. In this context, the most salient
feature of the AU Mic SED is the absence of emission excess at λ ≤ 25 µm (e.g., Liu et al.
2004; Chen et al. 2005). We assume the central peak represents the combined emission from
the star and dust, such that Fdust = Fcen − F∗ ≈ 0.25 mJy at 1.3 mm. Optically thin dust
emission at a temperature, T , has Fdust ≈ κνBν(T )Mdust/d2, where κν is the opacity spec-
trum, Bν the Planck function, Mdust the mass, and d = 9.91 pc. For a given dust population
characterized by κν, we computed the maximum T (and minimum Mdust) consistent with
-- 8 --
both the observed millimeter flux density and the infrared SED. We calculated various κν
for dust with the Weingartner & Draine (2001) "astrosilicate" composition and a power-law
size distribution n(a) ∝ a−3.5 between amin = 0.2 µm (the blow-out size; Strubbe & Chiang
2006) and amax values from 1 µm to 1 cm. For amax ≤ 100 µm, models of the central peak
over-predict the observed 60-70 µm emission if T > 35 K. However,
larger grains with
amax ≥ 1 mm at temperatures up to T ≈ 75 K can be accommodated without producing
an excess at λ ≤ 25 µm. These maximum T values are comparable to the expected dust
temperatures a few AU from the star, compatible with the emission size constraints (Rcen).
The corresponding minimum Mdust is ∼9× 1023 g, about 1% of the lunar mass. These calcu-
lations show that the central emission peak is consistent with a cool dust belt located .3 AU
from the central star, with a total mass comparable to the asteroid belt in our Solar System.
If this interpretation is correct, then ALMA observations at higher resolution can determine
its properties. Interestingly, the temperature of this putative inner belt is colder than the
∼ 190 K found to apply systematically to inner belts around F5-K0 stars by Morales et al.
(2011).
4.2. The Outer Dust Belt
Our modeling of the ALMA data locates the far edge of the outer emission belt with
high precision, ro = 40 AU, which matches closely the outer edge of the hypothesized "birth
ring" of colliding planetesimals. This analysis does not define the shape of the edge below the
∼ 6 AU resolution limit, but the truncation is reminiscent of the outer edge of the classical
Kuiper Belt (47 ± 1 AU; Trujillo & Brown 2001). The origins of such sharp edges remain
unclear: they could be from dynamical interactions (Ida et al. 2000; Boley et al. 2012), or
they may simply represent the initial conditions, where planetesimal formation was efficient
and successful in the primordial disk. Adopting the opacity used in §4.1 (κν = 2.7 cm2 g−1),
and assuming T ≈ 25 K (for 35-45 AU), the dust mass of this outer belt is 7×1025 g (consistent
with previous estimates; Liu et al. 2004), ∼100× more massive than the hypothesized inner
belt; the Kuiper Belt and asteroid belt have a similar mass ratio.
The mm-wave emission morphologies of cold belts of dusty debris reflect the dynamical
processes that shape the underlying planetesimal distributions. For AU Mic, our modeling
suggests that its outer emission belt can be described by an increasing emission profile with
a positive radial power-law index x ≈ 2.3±0.3. If we assume the emitting dust is in radiative
equilibrium with a temperature profile T ∝ r−0.5, this implies a rising surface density profile,
Σ ∝ r2.8, strongly peaked near 40 AU. A broad parent body ring with constant surface
density would produce a radial intensity profile with x ≈ −0.5, a value ruled out with
-- 9 --
high confidence (> 5σ). A rising behavior is predicted for "self-stirred" disks with ongoing
planet formation (Kenyon & Bromley 2002); in particular, the models of Kennedy & Wyatt
(2010) suggest Σ ∝ r7/3. However, the timescale required to assemble Pluto-sized bodies at
∼40 AU to initiate a collisional cascade around a low-mass star like AU Mic is much longer
than its ∼10 Myr age (Kenyon & Bromley 2008). Moreover, this scenario does not naturally
accommodate the presence of a separate, interior planetesimal belt. Of course, the still
modest resolution of the data is compatible with more complex scenarios, such as multiple
closely-spaced belts of different brightnesses that mimic a smooth gradient. Scattered light
observations of the AU Mic disk show asymmetries on both large and small scales, with
several peaks and depressions projected against the broad ansae in Figure 1, at radii beyond
the millimeter undulations (features A-E; see Fitzgerald et al. 2007). With such a steep
emission gradient in this outer belt, the data do not strongly constrain its width, or the
location of its inner edge. Our modeling indicates substantial emission from mm-sized grains
interior to 40 AU, in the ∼20-40 AU zone inferred to be highly depleted of µm-sized grains
from polarized scattered light (Graham et al. 2007).
The ALMA data show no clear evidence for asymmetries or substructure that would
signal planet-disk interactions. The hints of modulating millimeter brightness along the belt
in Figure 1 are insignificant in the residuals from subtracting a symmetric parametric model
(see Figure 2). This rules out substructure brighter than 90 µJy beam−1 (3 σ), corresponding
to dust clumps & 1% of the lunar mass (for the dust properties adopted above). Those
limits argue against over-densities of dust-producing planetesimals trapped in mean motion
resonances (Kuchner & Holman 2003), as might arise from the outward migration of planets
(Wyatt 2003). Given the young age of the system, the broad and smooth character of
the outer belt in the AU Mic disk may resemble the Kuiper Belt prior to the epoch of
Neptune's migration (Malhotra 1995). It is interesting that none of the claims of millimeter
emission clumps in debris disks have survived scrutiny at higher sensitivity (Pi´etu et al. 2011;
Hughes et al. 2011, 2012). It may be that any such features are effectively erased by collisions
(Kuchner & Stark 2010). We also find no significant centroid offset between the outer belt
and central peak, as might result from the secular perturbations of a planet in an eccentric
orbit (Wyatt et al. 1999). The limit on the displacement, ∆rcen < 1.9 AU (3 σ), corresponds
approximately to a limit on ae, where a is the semi-major axis and e is the eccentricity.
This limit can still accommodate a wide-orbit planet with modest eccentricity, similar to
Uranus. Such a planet could be responsible for stirring the disk to 40 AU in ∼10 Myr (e.g.
for a = 30 AU and e = 0.05, see eqn. 15 of Mustill & Wyatt 2009). Limits from high contrast
direct imaging admit Saturn-mass planets at these separations (Delorme et al. 2012).
-- 10 --
4.3. Concluding Remarks
The basic architecture of the AU Mic debris disk appears remarkably similar to the
Solar System, with a potential analog to the asteroid belt at a few AU, and a colder, more
massive, and apparently truncated counterpart of the Kuiper Belt extending to 40 AU. Future
observations are needed to determine if stellar processes could be responsible for emission
attributed to the asteroid belt, and to determine if the Solar System analogy extends to
include a planetary system like our own.
M.A.M. thanks NRAO for Student Observing Support funds. A.M.H. is supported by a
fellowship from the Miller Institute for Basic Research in Science. M.B. is funded through a
Space Science Enhancement Program grant from the Canadian Space Agency and an NSERC
Discovery Accelerator Supplement. E.C. acknowledges NSF grant AST-0909210. P.K. and
J.R.G. acknowledge support from NSF Award 0909188 and NASA Award NNX11AD21G
This paper makes use of the following ALMA data: ADS/JAO.ALMA#2011.0.00142.S.
ALMA is a partnership of ESO (representing its member states), NSF (USA) and NINS
(Japan), together with NRC (Canada) and NSC and ASIAA (Taiwan), in cooperation with
the Republic of Chile. The Joint ALMA Observatory is operated by ESO, AUI/NRAO and
NAOJ. The National Radio Astronomy Observatory is a facility of the National Science
Foundation operated under cooperative agreement by Associated Universities, Inc.
Facility: ALMA
-- 11 --
Table 1. ALMA Cycle 0 Observations of AU Mic
ID
SB-1
SB-2
SB-3
SB-4
Date (UT)
Antennas
PWV (mm)
2012 Apr 23 07:30 -- 09:26
2012 Apr 23 09:39 -- 11:03
2012 Apr 24 09:09 -- 11:19
2012 Jun 16 05:48 -- 08:02
17
16
18
20
1.7
1.7
3.0
0.7
Table 2. Model Parameters
Parameter
Description
Best-Fit
68% Confidence Interval
Fbelt
x
ri
ro
PA
αbelt
Fcen
∆rcen
σ2
cen
αcen
∆α
∆δ
belt flux density (mJy)
belt radial power law index
belt inner radius (AU)
belt outer radius (AU)
belt position angle (◦)
belt spectral index
Gaussian flux density (mJy)
Gaussian offset (AU)
Gaussian variance (AU2)
Gaussian spectral index
R.A. offset of belt center (′′)
Dec. offset of belt center (′′)
7.14
2.32
8.8
40.3
128.41
-0.15
0.32
0.71
≤5.9
-0.35
0.61
-0.03
+0.12, -0.25
+0.21, -0.31
+11.0, -1.0
+0.4, -0.4
+0.12, -0.13
+0.40, -0.58
+0.06, -0.06
+0.35, -0.51
(3 σ limit)
+2.1, -4.5
+0.02, -0.02
+0.02, -0.02
-- 12 --
REFERENCES
Augereau, J.-C., & Beust, H. 2006, A&A, 455, 987
Backman, D., Marengo, M., Stapelfeldt, K., et al. 2009, ApJ, 690, 1522
Boley, A. C., Payne, M. J., Corder, S., et al. 2012, ApJ, 750, L21
Bower, G. C., Bolatto, A., Ford, E. B., & Kalas, P. 2009, ApJ, 701, 1922
Chen, C. H., Patten, B. M., Werner, M. W., et al. 2005, ApJ, 634, 1372
Delorme, P., et al. 2012, A&A, 539, 72
Ertel, S., Wolf, S., & Rodmann, J. 2012, A&A, 544, A61
Fitzgerald, M. P., Kalas, P. G., Duchene, G., Pinte, C., & Graham, J. R. 2007, ApJ, 670,
536
Foreman-Mackey, D., Hogg, D. W., Lang, D., & Goodman, J. 2012, arXiv:1202.3665
Goodman, J., & Weare, J. 2010, Comm. App. Math. Comp. Sci., 5, 65
Graham, J. R., Kalas, P. G., & Matthews, B. C. 2007, ApJ, 654, 595
Hauschildt, P. H., Allard, F., & Baron, E. 1999, ApJ, 512, 377
Hughes, A. M., Wilner, D. J., Andrews, S. M., et al. 2011, ApJ, 740, 38
Hughes, A. M., Wilner, D. J., Mason, B., et al. 2012, ApJ, 750, 82
Ida, S., Larwood, J., & Burkert, A. 2000, ApJ, 528, 351
Kalas, P., Liu, M. C., & Matthews, B. C. 2004, Science, 303, 1990
Kennedy, G. M., & Wyatt, M. C. 2010, MNRAS, 405, 1253
Kenyon, S. J., & Bromley, B. C. 2002, ApJ, 577, L35
Kenyon, S. J., & Bromley, B. C. 2008, ApJS, 179, 451
Krist, J. E., Ardila, D. R., Golimowski, D. A., et al. 2005, AJ, 129, 1008
Kuchner, M. J., & Holman, M. J. 2003, ApJ, 588, 1110
Kuchner, M. J., & Stark, C. C. 2010, AJ, 140, 1007
-- 13 --
Kundu, M. R., Jackson, P. D., White, S. M., & Melozzi, M. 1987, ApJ, 312, 822
Leto, G., Pagano, I., Linsky, J. L., Rodon`o, M., & Umana, G. 2000, A&A, 359, 1035
Liu, M. C. 2004, Science, 305, 1442
Liu, M. C., Matthews, B. C., Williams, J. P., & Kalas, P. G. 2004, ApJ, 608, 526
Malhotra, R. 1995, AJ, 110, 420
Metchev, S. A., Eisner, J. A., Hillenbrand, L. A., & Wolf, S. 2005, ApJ, 622, 451
Morales, F. Y., Rieke, G. H., Werner, M. W., et al. 2011, ApJ, 730, L29
Mustill, A. J., & Wyatt, M. C. 2009, MNRAS, 399, 1403
Pi´etu, V., di Folco, E., Guilloteau, S., Gueth, F., & Cox, P. 2011, A&A, 531, L2
Su, K. Y. L., Rieke, G. H., Stapelfeldt, K. R., et al. 2009, ApJ, 705, 314
Strubbe, L. E., & Chiang, E. I. 2006, ApJ, 648, 652
Trujillo, C. A., & Brown, M. E. 2001, ApJ, 554, L95
van Leeuwen, F. 2007, A&A, 474, 653
Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
White, S. M., Lim, J., & Kundu, M. R. 1994, ApJ, 422, 293
Wilner, D. J., Andrews, S. M., & Hughes, A. M. 2011, ApJ, 727, L42
Wilner, D. J., Andrews, S. M., MacGregor, M. A., & Hughes, A. M. 2012, ApJ, 749, L27
Wyatt, M. C., Dermott, S. F., Telesco, C. M., et al. 1999, ApJ, 527, 918
Wyatt, M. C. 2003, ApJ, 598, 1321
Wyatt, M. C. 2006, ApJ, 639, 1153
Zuckerman, B., Song, I., Bessell, M. S., & Webb, R. A. 2001, ApJ, 562, L87
This preprint was prepared with the AAS LATEX macros v5.2.
|
1106.3926 | 1 | 1106 | 2011-06-20T14:46:26 | Rosetta-Alice Observations of Exospheric Hydrogen and Oxygen on Mars | [
"astro-ph.EP"
] | The European Space Agency's Rosetta spacecraft, en route to a 2014 encounter with comet 67P/Churyumov-Gerasimenko, made a gravity assist swing-by of Mars on 25 February 2007, closest approach being at 01:54UT. The Alice instrument on board Rosetta, a lightweight far-ultraviolet imaging spectrograph optimized for in situ cometary spectroscopy in the 750-2000 A spectral band, was used to study the daytime Mars upper atmosphere including emissions from exospheric hydrogen and oxygen. Offset pointing, obtained five hours before closest approach, enabled us to detect and map the HI Lyman-alpha and Lyman-beta emissions from exospheric hydrogen out beyond 30,000 km from the planet's center. These data are fit with a Chamberlain exospheric model from which we derive the hydrogen density at the 200 km exobase and the H escape flux. The results are comparable to those found from the the Ultraviolet Spectrometer experiment on the Mariner 6 and 7 fly-bys of Mars in 1969. Atomic oxygen emission at 1304 A is detected at altitudes of 400 to 1000 km above the limb during limb scans shortly after closest approach. However, the derived oxygen scale height is not consistent with recent models of oxygen escape based on the production of suprathermal oxygen atoms by the dissociative recombination of O2+. | astro-ph.EP | astro-ph |
Rosetta-Alice Observations of Exospheric Hydrogen
and Oxygen on Mars
Paul D. Feldmana,∗, Andrew J. Stefflb, Joel Wm. Parkerb, Michael F. A'Hearnc,
Jean-Loup Bertauxd, S. Alan Sternb, Harold A. Weavere, David C. Slaterf,
Maarten Versteegf, Henry B. Throopb, Nathaniel J. Cunninghamg, Lori M.
Feagac
aJohns Hopkins University, Department of Physics and Astronomy, 3100 N. Charles Street,
bSouthwest Research Institute, Department of Space Studies, Suite 300, 1050 Walnut Street,
Baltimore, MD 21218 USA
Boulder CO 80302-5150 USA
cDepartment of Astronomy, University of Maryland, College Park MD 20742-2421 USA
dLATMOS, CNRS/UVSQ/IPSL, 11 Boulevard d'Alembert, 78280 Guyancourt, France
eJohns Hopkins University Applied Physics Laboratory, Space Department, 11100 Johns
Hopkins Road, Laurel, MD 20723-6099 USA
fSouthwest Research Institute, P. O. Drawer 28510, San Antonio TX 78228-0510 USA
gPhysics Department, Nebraska Wesleyan University, Lincoln, NE 68504-2794 USA
Abstract
The European Space Agency's Rosetta spacecraft, en route to a 2014 en-
counter with comet 67P/Churyumov-Gerasimenko, made a gravity assist swing-
by of Mars on 25 February 2007, closest approach being at 01:54 UT. The Alice
instrument on board Rosetta, a lightweight far-ultraviolet imaging spectrograph
optimized for in situ cometary spectroscopy in the 750 -- 2000 A spectral band, was
used to study the daytime Mars upper atmosphere including emissions from exo-
spheric hydrogen and oxygen. Offset pointing, obtained five hours before closest
approach, enabled us to detect and map the HI Lyman-a and Lyman-b emissions
from exospheric hydrogen out beyond 30,000 km from the planet's center. These
data are fit with a Chamberlain exospheric model from which we derive the hy-
drogen density at the 200 km exobase and the H escape flux. The results are
comparable to those found from the the Ultraviolet Spectrometer experiment on
the Mariner 6 and 7 fly-bys of Mars in 1969. Atomic oxygen emission at 1304 A
is detected at altitudes of 400 to 1000 km above the limb during limb scans shortly
after closest approach. However, the derived oxygen scale height is not consistent
with recent models of oxygen escape based on the production of suprathermal
1
2 .
oxygen atoms by the dissociative recombination of O+
Keywords: Mars, Mars atmosphere, Atmospheres, evolution
1. Introduction
The extended atomic hydrogen corona of Mars was first detected by the ultra-
violet spectrometer experiments on the Mariner 6 and 7 spacecraft that measured
resonantly scattered solar Lyman-a
radiation (Barth et al., 1971) to a planetocen-
tric distance of 24,000 km. This was followed by a similar experiment on the or-
biting Mariner 9 mission (Barth et al., 1972). Anderson and Hord (1971), using
radiative transfer theory, analyzed the early data to derive a hydrogen escape rate,
which they found to be compatible with the water photodissociation rate at Mars.
Barth et al. (1972) recognized that the apparently constant H escape rate derived
from Mariner 6, 7, and 9, if extended backward over geological time scales, would
result in an oxygen abundance in the lower atmosphere several orders of magni-
tude larger than observed. McElroy (1972) suggested that dissociative recombi-
nation of O+
2 , the dominant ion in the atmosphere, would produce oxygen atoms
with sufficient energy to escape. However, since the energetic oxygen atoms are
mostly produced below the exobase, determination of the escaping fraction re-
quires detailed modeling, represented by the recent work of Lammer et al. (2003),
Fox and Ha´c (2009), Shematovich et al. (2007), Valeille et al. (2009, 2010), and
others. Barth et al. (1971) also reported observations of O I l 1304 emission up
to 700 km. These data, together with those from Mariner 9, were interpreted by
Strickland et al. (1973) in terms of a cool, optically thick oxygen exosphere. Since
then, the only measurement of exospheric oxygen on Mars is from the SPICAM
instrument on Mars Express (Chaufray et al., 2009), but those data extended only
up to 400 km where radiative transfer effects in the O I l 1304 multiplet are sig-
nificant and the evidence for a hot component of O atoms was inconclusive.
Until recently there have also been no additional measurements of the ex-
tended hydrogen atmosphere on Mars. Chaufray et al. (2008), again with SPI-
CAM, have measured H I Lyman-a emission at 1216 A, but only up to 4,000 km.
Again, they find that a two temperature H exosphere is possible, although the
∗Corresponding author
Email address: [email protected] (Paul D. Feldman)
Preprint submitted to Icarus
November 8, 2018
result is inconclusive. Clarke et al. (2009) also suggest a two-component H distri-
bution from monochromatic Lyman-a
images of Mars showing the H corona out
to 4 Mars radii (RM) taken by the Solar Blind Channel of the Advanced Camera
for Surveys on HST. Radiative transfer modeling is necessary to extract densities
from both of these data sets. Clarke et al. also noted a variation in the Lyman-a
brightness over a period of two months which they ascribe to a seasonal variation
in the H2O loss rate.
We report here on observations of both the extended hydrogen and oxygen
coronae of Mars made with the Alice far-ultraviolet imaging spectrograph (Stern et al.,
2007) on Rosetta during the spacecraft's gravity assist swing-by of Mars on 25
February 2007. Offset exposures enabled us to detect and map the H I Lyman-a
and Lyman-b emissions to beyond 30,000 km from the planet's center. Moreover,
except near the planet's limb, the Lyman-b emission is optically thin, allowing us
to use a spherical Chamberlain model to determine the temperature and density
of H at the exobase without the need for radiative transfer modeling. From limb
pointings at spacecraft distances closer to the planet, oxygen emission above the
exobase is detected, allowing us to constrain the density of hot oxygen without
the need for radiative transfer modeling. These data can be used to derive atomic
escape rates and address the question of stoichiometric loss of water vapor from
Mars.
2. Observations
Rosetta approached Mars from the day side making its closest approach (CA)
at 01:54 UT on 25 February 2007 at an altitude of 250 km. In order to satisfy
the scientific goals of the various remote sensing instruments (see, e.g., Coradini,
2010), the common instrument boresight was programmed for a number of fixed
pointings towards both the sunlit and dark hemispheres of Mars and offset from
Mars, as well as raster scans across the sunlit limb. Due to operational constraints,
observations were not possible during the immediate CA period. The pointings of
interest in this paper (denoted by the operational designation ALxx) were AL03
pre-CA, centered on the sunlit disk, AL10E, offset pointings centered 2.5◦ and
7.5◦ from Mars along the equator, and AL11B, scans across the illuminated cres-
cent post-CA. Observation start times and geometry parameters are given in Ta-
ble 1. At the time of closest approach, Mars was 1.445 AU from the Sun and the
areocentric longitude, Ls, was 189.9◦. Solar activity was very low for an extended
time, including when Earth faced the same solar longitude 9 days earlier, with
F10.7 ≈72 at 1 AU.
3
Alice is a lightweight, low-power, imaging spectrograph optimized for in situ
cometary far-ultraviolet (FUV) spectroscopy. It is designed to obtain spatially-
resolved spectra in the 750-2000 A spectral band with a spectral resolution be-
tween 8 and and 12 A for extended sources that fill its field-of-view. The slit is in
the shape of a dog bone, 5.5◦ long, with a width of 0.05◦ in the central 2.0◦ while
the ends are 0.10◦ wide. Each spatial pixel along the slit is 0.30◦. Alice employs
an off-axis telescope feeding a 0.15-m normal incidence Rowland circle spectro-
graph with a concave holographic reflection grating. The imaging microchannel
plate detector utilizes dual solar-blind opaque photocathodes (KBr and CsI) and
employs a two-dimensional delay-line readout. Details of the instrument are given
by Stern et al. (2007).
The Alice slit geometry, illustrating the shape of the multi-segment slit, is
shown in Fig. 1 for the first pre-CA offset pointing of 2.5◦. The second offset
moved the slit an additional 5.0◦ away from Mars parallel to the Martian equator.
Except near the limb, the only features seen in the offset spectra are H I Lyman-
, and these data are used to extract the spatial profiles of these
and Lyman-b
emissions.
A similar diagram for the post-CA limb scans, beginning 25 February 2007
at UT 03:33:02, is shown in Fig. 2. At this time each spatial pixel projected to
280 km in altitude but as Rosetta receded from Mars the projected size of each
pixel increased. The scan slowly shifted the boresight ∼250 km towards Mars
over a 15-minute period. This is illustrated in Fig. 3. Because of the scanning
motion, these spectra were acquired in "pixel-list" mode, that is the position of
each photon count is recorded together with a time tag so that spectra could be
reconstructed with the motion accounted for. Because the Rosetta-Alice instru-
ment has only a single data buffer, the data gaps seen in Fig. 3 result from the
time required to read out the buffer to the spacecraft. The typical time to fill the
buffer was 30 s, so in practice we accumulated individual spectra corresponding
to ∼30 s integrations. Even so, to detect O I emission at high altitudes, we need
to co-add multiple spectra, as described below.
3. Data Analysis
3.1. Calibration
During the gravity assist swing-bys of both Mars and the Earth, the Rosetta
instruments were powered on and operated primarily to provide flight verifica-
tion of instrument performance and to acquire calibration data such as standard
ultraviolet star fluxes and detector flat-fields. There were also opportunities to
4
a
exercise the full range of instrument parameters that could be adjusted by remote
command in flight in order to optimize the signal-to-noise performance of the
instrument. For the Mars swing-by, the detector high voltage level was set at
-- 3.8 kV. Subsequent operations and analysis showed the optimum setting to be
-- 3.9 kV and all observations beginning in the fall of 2007 were made at that volt-
age. Nevertheless, by comparison of stellar standards at different voltages (and at
different times), we are able to transfer the current absolute flux calibration to the
epoch of the Mars swing-by and this is incorporated into the current version of
the data pipeline (version 3) with which all of the data have since been processed.
The data used in this study are publicly available, both from NASA's Planetary
Data System (http://pdssbn.astro.umd.edu/) and ESA's Planetary Science
Archive (http://www.rssd.esa.int/?project=PSA).
To derive spatial profiles of the observed emissions requires an accurate flat-
field calibration. This poses a rather acute problem for Lyman-a
at 1216 A be-
cause the instrument was designed for this wavelength to fall in the gap between
the KBr and CsI photocathode coatings. The intent was to utilize the low detection
efficiency of the bare microchannel plate to compensate for the very high expected
Lyman-a photon flux from the comet. However, because of a slight misalignment
in the coating edges, the Lyman-a
sensitivity varies by a factor of two along the
portion of the detector that is mapped onto the sky. In contrast, at Lyman-b
the
variation across the slit is ∼1.4. There is also a ±20% odd-even detector row
effect that is accentuated at the lower detector high voltage of -- 3.8 kV.
Fortuitously, following the Mars encounter, Rosetta-Alice was pointed to-
wards Jupiter and recorded many hours of spectra in support of the New Hori-
zons fly-by of Jupiter in February 2007. These observations were made with the
same instrument parameters used for the Mars observations. Since, from the or-
bit of Mars the Jovian system only filled a single row of the detector, high S/N
measurements of detector flat-fields at Lyman-a and Lyman-b were obtained, to-
gether with a measure of grating scattered Lyman-a as a function of detector row.
An added bonus is that since Jupiter was only 20◦ away on the sky from the
coordinates of the offset pointing, these observations provided a measure of the
interplanetary Lyman-a and Lyman-b background for the Mars observations. We
used a co-added accumulation of 630,000 s of data obtained between 1 March and
10 March 2007 to derive the flat-fields used in the analysis described below.
3.2. Mars dayglow spectrum
We briefly discuss the dayglow spectrum of Mars, obtained under AL03 pre-
CA. It consisted of four exposures, each of 1028 seconds. A composite of the
5
five central rows of the sum of four exposures is shown in Fig. 4. The viewing
parameters are given in Table 1. At the start of the sequence, the angular diameter
of Mars was 1.62◦ so that the five central rows of the detector were uniformly
filled with the illuminated Martian disk. For comparison with previous work, we
also show the Mars full disk spectrum recorded by the Hopkins Ultraviolet Tele-
scope (HUT) on board the Space Shuttle in March 1995 (a full solar cycle earlier)
(Feldman et al., 2000), convolved to the spectral resolution of Alice. At the time,
Mars was 1.666 AU from the Sun, Ls was 70.5◦, and F10.7 was ≈75. The HUT
spectrum is multiplied by a factor of 0.80 to match the brightness observed by
Alice, and this is quite good agreement considering the Alice calibration uncer-
tainty and the fact that the HUT spectrum measured the integrated disk brightness,
not just a central stripe of the disk. The comparison also serves to validate the
wavelength calibration and provides a reference spectrum with which to compare
the exospheric spectra obtained in AL10E and AL11B. As noted by Barth et al.
(1972) and Leblanc et al. (2006), with the exception of the H I Lyman series and
O I l 1304, Mars' dayglow emissions, principally CO and C I, are confined to alti-
tudes below ∼200 km.
3.3. Offset exposures
The two offset exposures were taken immediately following the full disk spec-
tra as Rosetta approached Mars. H I Lyman-b was detected along the full length
of the Alice slit for both offsets. Because the slit is segmented, for each row along
the slit the observed Lyman-b
signal was fit to a gaussian profile superimposed
on a grating-scattered background that was fit to a second-order polynomial. The
flux was then obtained by integrating under the gaussian and then correcting for
the detector flat-field that was derived from subsequent Jupiter observations as de-
scribed in Section 3.1. The result is shown as a histogram in Fig. 5. The errors
shown are statistical in the count rate. The Lyman-a profile is similarly derived
except that the detector background, due to dark counts, was negligible, and is
shown in Fig. 6. The shape of this profile is very similar to the slant intensity
profiles derived from the Mariner 6 and 7 fly-bys by Barth et al. (1971), although
lower in absolute brightness as the Mariner fly-bys occurred at a time of high
solar activity. Neither Fig. 5 nor Fig. 6 have had the interplanetary background
subtracted as did the plots of Barth et al. The model fits to the data are discussed
below in Section 4.1.
6
3.4. Limb scan spectra
From the data acquired in pixel list mode during the limb scans schematically
illustrated in Fig. 3, we can extract a spectrum spanning a given time interval.
However, because the line-of-sight of the spectrogram was changing its position
above the limb with time, it is necessary to balance the motion with the need
for a sufficiently long integration time to obtain an adequate signal-to-noise ratio
for weak emission features. At the same time, we need to avoid contamination
of the spectrum by thermospheric emissions. This was done by co-adding the
photon counts from the first four exposures for row 14; the first 12 exposures for
row 15; and the last 14 exposures for rows 16 and 17. This results in an altitude
weighted average with a trapezoidal shape of ≈320 km centered at 420, 665, 910,
and 1240 km, respectively for rows 14 to 17, respectively.
Examples of extracted spectra for rows 15 and 16 are shown in Fig. 7. Note
that only H I and O I emissions are detected. A possible feature at 1657 A in the
row 15 spectrum that could be a signature of escaping carbon atoms (Fox and Ha´c,
1999; Cipriani et al., 2007), is most likely an instrumental artifact as no emission
at this wavelength appears in the row 14 spectrum. The background is due to
grating scattered Lyman-a
, which is variable from row to row. Also, the spectra
have not been corrected for the odd-even row variation noted in Section 3.1 which
is estimated to be ∼25% for Lyman-b and ∼10% for O I l 1304, based on the disk
observations discussed in Section 3.2.
4. Discussion
4.1. Exospheric hydrogen model
For the analysis of the Lyman-b profile we follow the same procedure as
Anderson and Hord (1971), using the exospheric model of Chamberlain (1963)
but ignoring satellite orbits, as they can be excluded by the observed Lyman-b
brightness near 30,000 km planetocentric distance. For solar minimum conditions
we take the exobase to be at 200 km and the exobase temperature, Te, to be 200 K
(Krasnopolsky, 2002; Fox and Ha´c, 2009), leaving the hydrogen density at this
level as a variable. We assume that Lyman-b
is optically thin and calculate a flu-
orescence efficiency (g-factor) of 3.9 × 10−6 photons s−1 atom−1 at 1 AU using
a solar minimum line profile and flux from Lemaire et al. (2002). The interplane-
tary Lyman-b background is fixed at 1.0 rayleigh based on the subsequent Jupiter
observations that were used to derive the detector flat-field at 1026 A (see Sec-
tion 3.1). All fluxes are referenced to row 15 which is the nominal Alice boresight
and which is used for almost all of the stellar calibration measurements.
7
The result is shown by the solid curve in Fig. 5. The derived H density at
200 km is 2.5 × 105 cm−3 and the escape flux is 7.8 × 107 cm−2 s−1. The model
atmosphere is given in Table 2. Optical depth unity along the line-of-sight is at
∼5,000 km projected planetocentric distance. From the radiative transfer model
of Anderson and Hord (1971), applied to the Mariner Lyman-a data, we expect
the single scattering intensity to be about a factor of two higher than the radiative
transfer corrected intensity at this optical depth, and this is consistent with the
data shown in Fig. 5. Anderson and Hord, for their solar maximum observations,
assuming Te = 350 K at an exobase altitude of 250 km, found an H density and
escape flux of 3.0 × 104 cm−3 and 1.8 × 108 cm−2 s−1, respectively. Fig. 5 also
shows a model for 260 K (dashed line), which, normalized at 30,000 km, provides
what superficially appears to be a better fit to the observed profile. While it is
difficult to choose between the models for planetocentric distances greater than
10,000 km, the absence of a strong optical depth effect in the latter suggests that
the lower temperature model, corresponding to a typical exospheric temperature
at solar minimum, is probably correct. For 260 K, the H density and escape flux
are 9.0 × 104 cm−3 and 1.34 × 108 cm−2 s−1, respectively.
The same models, applied to Lyman-a using a Lyman-a
ratio of 250
derived from the IPM measurements, are shown in Fig. 6. The agreement is excel-
lent and the deviation from the optically thin emission is consistent with an optical
depth along the line-of-sight of 1 at 10,000 km planetocentric distance. There is
no apparent need for a suprathermal H component as suggested by Chaufray et al.
(2008) and Clarke et al. (2009).
/Lyman-b
An interesting measurement of exospheric Lyman-a emission from Mars Ex-
press has recently been reported by Galli et al. (2006). They found that the Neutral
Particle Detector of the ASPERA-3 experiment was sensitive to Lyman-a photons
and measured a signal, attributed to exospheric hydrogen, out to a tangent height
of 7,250 km above the Martian limb. Considering their large measurement uncer-
tainties that include calibration, statistics, pointing, and background subtraction,
their measured emission profile is in general accord with the Alice data shown in
Fig. 6. However, they interpret this profile in terms of an optically thin resonance
scattering model and derive an apparent temperature > 600K. As noted above, we
find that the Lyman-a emission is optically thick below 10,000 km planetocentric
distance (∼6,600 km above the limb), which leads to a flatter spatial distribution
and consequently the appearance of a higher than actual exospheric temperature.
8
4.2. Two-component oxygen model
For oxygen we use a two-component model, again taking for the cold com-
ponent, Te, to be 200 K, and an oxygen density at 200 km of 3.0 × 107 cm−3
(Fox and Ha´c, 2009). The curve in Fig. 8 represents an added hot component of
1200 K with an oxygen density at 200 km of 1.0 × 105 cm−3 (see Table 2). The
density decrease with altitude is considerably faster than the predictions of recent
exospheric models of Chaufray et al. (2009) and Valeille et al. (2010) based on
a hot atomic O source due to dissociative recombination of O+
2 , and which are
necessary to support a stoichiometric escape of water vapor from the atmosphere
of Mars. However, it has been noted that such inferences from a single viewing
geometry during a period of low solar activity can be misleading as this mecha-
nism is quite sensitive to solar activity and is dependent on solar zenith angle at
the observation point.
Nevertheless, the present observations raise concern about some of the as-
sumptions and physical parameters used in the recent modeling of oxygen escape
from Mars. Fox and Ha´c (2009), in their comparison of exobase and Monte Carlo
models, summarize the literature on modeling efforts from the past few decades
and conclude that "efforts to balance the escape rates in the stoichiometric propor-
tion of water are premature." Fox and Ha´c focus on a comparison of escape rates
and therefore do not compute the oxygen density profiles from their models. Such
calculation is warranted by the present data which would allow for a determination
of the line-of-sight column densities appropriate to the Alice observations.
Similar questions arise in the analogous modeling of the hot oxygen environ-
ment around Venus (Groller et al., 2010). Bovino et al. (2011) discuss the need
for accurate data on energy transfer collisions between hot oxygen atoms and the
neutral atmosphere (they are mainly interested in helium) which is at the core of
the escape models. Finally, we note that Simon et al. (2009) point out that an ad-
ditional constraint on the models might be provided by SPICAM measurements
of the forbidden O I l 2972 (1S -- 3P) line, which is produced by both photodisso-
2 .
ciation of CO2 and by dissociative recombination of O+
5. Conclusion
The Rosetta swing-by of Mars on 25 February 2007, provided the first spec-
troscopic observations of exospheric hydrogen and oxygen on Mars from outside
Mars' atmosphere since the Mariner 6 and 7 fly-bys in 1969. The spatial distribu-
tion of H I Lyman-a out to beyond 30,000 km from the planet's center is similar to
9
that found from Mariner 6 and 7. A Chamberlain model, with current solar mini-
mum model values of exospheric temperature of 200 K and an exobase altitude of
200 km, provides a good fit to the observed Lyman-b profile, although the data do
not exclude temperatures up to ∼260 K. A suprathermal component, suggested by
several authors, is not needed to match the data. The hydrogen escape flux derived
from the 200 K model, 7.8 ×107 cm−2 s−1, is comparable to that derived from the
earlier measurements. The distribution of atomic oxygen, derived from 1304 A
emission observed to altitudes of 1000 km above the limb, is not consistent with
recent models of oxygen escape based on the production of suprathermal oxygen
atoms by dissociative recombination of O+
2 .
Acknowledgments
We thank the ESA Rosetta Science Operations Centre (RSOC) and Mission
Operations Center (RMOC) teams for their expert and dedicated help in planning
and executing the Alice observations of Mars. We thank Darrell Strobel for help-
ful discussions. The Alice team acknowledges continuing support from NASA's
Jet Propulsion Laboratory through contract 1336850 to the Southwest Research
Institute. The work at Johns Hopkins University was supported by a sub-contract
from Southwest Research Institute.
10
References
Anderson, Jr., D.E., Hord, C.W., 1971. Mariner 6 and 7 ultraviolet spectrometer
J. Geophys. Res. 76,
experiment: Analysis of hydrogen Lyman-alpha data.
6666 -- 6673.
Barth, C.A., Hord, C.W., Pearce, J.B., Kelly, K.K., Anderson, G.P., Stewart, A.I.,
1971. Mariner 6 and 7 ultraviolet spectrometer experiment: Upper atmosphere
data. J. Geophys. Res. 76, 2213 -- 2227.
Barth, C.A., Stewart, A.I., Hord, C.W., Lane, A.L., 1972. Mariner 9 Ultravi-
olet Spectrometer Experiment: Mars Airglow Spectroscopy and Variations in
Lyman Alpha. Icarus 17, 457 -- 468.
Bovino, S., Zhang, P., Gianturco, F.A., Dalgarno, A., Kharchenko, V., 2011. En-
ergy transfer in O collisions with He isotopes and Helium escape from Mars.
Geophys. Res. Lett. 38, 2203.
Chamberlain, J.W., 1963. Planetary coronae and atmospheric evaporation. Planet.
Space Sci. 11, 901 -- 960.
Chaufray, J.Y., Bertaux, J.L., Leblanc, F., Qu´emerais, E., 2008. Observation of
the hydrogen corona with SPICAM on Mars Express. Icarus 195, 598 -- 613.
Chaufray, J.Y., Leblanc, F., Qu´emerais, E., Bertaux, J.L., 2009. Martian oxy-
gen density at the exobase deduced from O I 130.4-nm observations by Spec-
troscopy for the Investigation of the Characteristics of the Atmosphere of Mars
on Mars Express. J. Geophys. Res. 114, 2006.
Cipriani, F., Leblanc, F., Berthelier, J.J., 2007. Martian corona: Nonthermal
sources of hot heavy species. J. Geophys. Res. 112, 7001.
Clarke, J.T., Bertaux, J., Chaufray, J., Gladstone, R., Quemerais, E., Wilson, J.K.,
2009. HST Observations Of The Extended Hydrogen Corona Of Mars. Bull.
Am. Astron. Soc. 41, 49.11 (abstract).
Coradini, A., et al., 2010. Martian atmosphere as observed by VIRTIS-M on
Rosetta spacecraft. J. Geophys. Res. 115, 4004.
Feldman, P.D., Burgh, E.B., Durrance, S.T., Davidsen, A.F., 2000. Far-Ultraviolet
Spectroscopy of Venus and Mars at 4 A Resolution with the Hopkins Ultraviolet
Telescope on Astro-2. Astrophys. J. 538, 395 -- 400.
11
Fox, J.L., Ha´c, A., 1999. Velocity distributions of C atoms in CO+ dissocia-
tive recombination: Implications for photochemical escape of C from Mars. J.
Geophys. Res. 104, 24729 -- 24738.
Fox, J.L., Ha´c, A.B., 2009. Photochemical escape of oxygen from Mars: A com-
parison of the exobase approximation to a Monte Carlo method. Icarus 204,
527 -- 544.
Galli, A., Wurz, P., Lammer, H., Lichtenegger, H.I.M., Lundin, R., Barabash, S.,
Grigoriev, A., Holmstrom, M., Gunell, H., 2006. The Hydrogen Exospheric
Density Profile Measured with ASPERA-3/NPD. Space Sci. Rev. 126, 447 --
467.
Groller, H., Shematovich, V.I., Lichtenegger, H.I.M., Lammer, H., Pfleger, M.,
Kulikov, Y., Macher, W., Amerstorfer, U.V., Biernat, H.K., 2010. Venus' atomic
hot oxygen environment. J. Geophys. Res. 115, 12017.
Krasnopolsky, V.A., 2002. Mars' upper atmosphere and ionosphere at low,
medium, and high solar activities: Implications for evolution of water. J. Geo-
phys. Res. 107, 5128.
Lammer, H., Lichtenegger, H.I.M., Kolb, C., Ribas, I., Guinan, E.F., Abart, R.,
Bauer, S.J., 2003. Loss of water from Mars:Implications for the oxidation of
the soil. Icarus 165, 9 -- 25.
Leblanc, F., Chaufray, J.Y., Lilensten, J., Witasse, O., Bertaux, J., 2006. Mar-
tian dayglow as seen by the SPICAM UV spectrograph on Mars Express. J.
Geophys. Res. 111, E09S11.
Lemaire, P., Emerich, C., Vial, J.C., Curdt, W., Schuhle, U., Wilhelm, K., 2002.
Variation of the full Sun hydrogen Lyman-a
and b profiles with the activity
cycle, in: Wilson, A. (Ed.), From Solar Min to Max: Half a Solar Cycle with
SOHO, Noordwijk: ESA SP-508. pp. 219 -- 222.
McElroy, M.B., 1972. Mars: An Evolving Atmosphere. Science 175, 443 -- 445.
Shematovich, V.I., Tsvetkov, G.A., Krestyanikova, M.A., Marov, M.Y., 2007.
Stochastic models of hot planetary and satellite coronas: Total water loss in the
Martian atmosphere. Solar System Res. 41, 103 -- 108.
12
Simon, C., Witasse, O., Leblanc, F., Gronoff, G., Bertaux, J., 2009. Dayglow on
Mars: Kinetic modelling with SPICAM UV limb data. Planet. Space Sci. 57,
1008 -- 1021.
Stern, S.A., Slater, D.C., Scherrer, J., Stone, J., Versteeg, M., A'Hearn, M.F.,
Bertaux, J.L., Feldman, P.D., Festou, M.C., Parker, J.W., Siegmund, O.H.W.,
2007. Alice: The Rosetta Ultraviolet Imaging Spectrograph. Space Sci. Rev.
128, 507 -- 527.
Strickland, D.J., Stewart, A.I., Barth, C.A., Hord, C.W., Lane, A.L., 1973.
Mariner 9 ultraviolet spectrometer experiment: Mars atomic oxygen 1304- A
emission. J. Geophys. Res. 78, 4547 -- 4559.
Valeille, A., Combi, M.R., Tenishev, V., Bougher, S.W., Nagy, A.F., 2010. A
study of suprathermal oxygen atoms in Mars upper thermosphere and exosphere
over the range of limiting conditions. Icarus 206, 18 -- 27.
Valeille, A., Tenishev, V., Bougher, S.W., Combi, M.R., Nagy, A.F., 2009. Three-
dimensional study of Mars upper thermosphere/ionosphere and hot oxygen
corona: 1. General description and results at equinox for solar low conditions.
J. Geophys. Res. 114, 11005.
13
Table 1: Observation Parameters.
Observation ID
Start time (UT 2007)
Boresight pointing
Distance to Marsa (km)
Solar elongationa
Longitude of tangent point at limbb
Latitude of tangent point at limbb
Solar zenith angle at tangent pointb
Data mode
Total integration time (s)
aAt start of observation.
bAt sub-observer point for AL03.
AL03
AL10E
AL11B
24 Feb 18:28:14
disk center
239,800
164.3◦
213.9◦
-- 1.8◦
15.7◦
Histogram
4112
24 Feb 20:13:14
offset 2.5◦, 7.5◦ W
25 Feb 03:33:02
E limb scan
184,200
164.0◦
149.7◦
-- 4.1◦
74.3◦
Histogram
483 + 1028
51,600
24.3◦
263.6◦
-- 26.5◦
67.9◦
Pixel list
14
Table 2: Model Atmosphere Densities.
O (cold)
(cm−3)
3.00 × 107
6.86 × 104
288.0
2.04
0.023
3.7 × 10−4
O (hot)
(cm−3)
1.00 × 105
3.70 × 104
1.49 × 104
6520.
3080.
1550.
830.
Altitude
(km)
200
400
600
800
1000
1200
1400
1600
1800
2000
4000
6000
8000
10000
15000
20000
25000
30000
35000
H
(cm−3)
2.50 × 105
1.69 × 105
1.19 × 105
8.61 × 104
6.42 × 104
4.90 × 104
3.82 × 104
3.03 × 104
2.44 × 104
1.99 × 104
4330.
1590.
757.
422.
144.
67.5
37.5
23.2
15.5
15
FIGURE CAPTIONS
Fig. 1. Projection of the Alice slit on the sky with the common boresight offset
2.5◦ from the center of Mars along the equator. Orange grid lines outline the
illuminated region of the disk. The shape of the multi-segment slit is shown.
Detector row numbers increase from left to right with the boresight (+) in row 15.
Fig. 2. Projection of the Alice slit on Mars at the beginning of the limb scans
following closest approach. Orange grid lines outline the illuminated crescent of
the disk. Detector row numbers increase from right to left with the boresight (+)
in row 15.
Fig. 3. Projection of individual rows of the Alice slit above the Mars limb dur-
ing the slow limb scan following closest approach. Initially, each spatial pixel
projected to 280 km in altitude but as Rosetta receded from Mars the projected
size of each pixel increased. The scan also slowly shifted the boresight ∼250 km
towards Mars over a 15-minute period. The cross-hatched areas indicate the time
during which photon events were accumulated while the horizontal lines show the
times for which photon events were co-added for each row.
Fig. 4. The dayglow spectrum of Mars is a composite of the five central rows
of four exposures, each of 1028 seconds. For comparison with previous work,
we also show the Mars full disk spectrum recorded by the Hopkins Ultraviolet
Telescope (HUT) at 4 A spectral resolution in March 1995 (a full solar cycle
earlier) (Feldman et al., 2000), convolved to the spectral resolution of Alice and
multiplied by a factor of 0.8.
Fig. 5. H I Lyman-b was detected along the full length of the Alice slit for both
offsets. The data are shown as a histogram: black, 2.5◦ offset, red: 7.5◦ offset.
The errors shown are statistical in the count rate. Optically thin Chamberlain
models without satellite orbits for T=200 K (solid line) and 260 K (dashed line)
are shown, superimposed on a 1 rayleigh interplanetary background. The radius
of Mars is indicated.
Fig. 6. Same as Fig. 5 for H I Lyman-a
. The shape of this profile is very similar to
the slant intensity profiles derived from the Mariner 6 and 7 fly-bys by Barth et al.
(1971), although lower in absolute brightness as the Mariner fly-bys occurred at
a time of high solar activity. The same models are shown, superimposed on an
16
interplanetary background of 250 rayleighs.
Fig. 7. Extracted limb scan spectra. For row 15 (top), the photon counts from the
first 12 exposures (376 s, see Fig. 3) were co-added. For row 16 (bottom), the last
14 exposures (476 s) were co-added. Only H I and O I emissions are detected.
Fig. 8. Extracted O I l 1304 brightness from rows 14 -- 17 as a function of altitude
above the Martian limb. The vertical bars illustrate the extent of the trapezoidal
altitude weighting function for each row while the horizontal bars are the statis-
tical uncertainty in the count rate. The two-component oxygen model shown is
described in the text.
17
Figure 1: Projection of the Alice slit on the sky with the common boresight offset 2.5◦ from the
center of Mars along the equator. Orange grid lines outline the illuminated region of the disk. The
shape of the multi-segment slit is shown. Detector row numbers increase from left to right with
the boresight (+) in row 15.
18
Figure 2: Projection of the Alice slit on Mars at the beginning of the limb scans following closest
approach. Orange grid lines outline the illuminated crescent of the disk. Detector row numbers
increase from right to left with the boresight (+) in row 15.
19
Row 16
Row 15
Row 14
Row 13
1200
1000
800
600
400
200
0
)
m
k
(
b
m
i
l
s
r
a
M
e
v
o
b
a
e
c
n
a
t
s
D
i
−200
30
35
40
Time from 2007−02−25T03:00 (minutes)
45
50
Figure 3: Projection of individual rows of the Alice slit above the Mars limb during the slow limb
scan following closest approach. Initially, each spatial pixel projected to 280 km in altitude but as
Rosetta receded from Mars the projected size of each pixel increased. The scan also slowly shifted
the boresight ∼250 km towards Mars over a 15-minute period. The cross-hatched areas indicate
the time during which photon events were accumulated while the horizontal lines show the times
for which photon events were co-added for each row.
20
CI
CI
CO A−X
OII
OI
OI
OI
OI
Lyg
Lyb +OI
CO CO
ArI
10
8
6
4
2
0
)
1
−
Å
s
h
g
e
y
a
R
l
i
(
s
s
e
n
t
h
g
i
r
B
800
1000
1200
Wavelength (Å)
1400
1600
1800
Figure 4: The dayglow spectrum of Mars is a composite of the five central rows of four exposures,
each of 1028 seconds. For comparison with previous work, we also show the Mars full disk
spectrum recorded by the Hopkins Ultraviolet Telescope (HUT) at 4 A spectral resolution in March
1995 (a full solar cycle earlier) (Feldman et al., 2000), convolved to the spectral resolution of Alice
and multiplied by a factor of 0.8.
21
Lyman−b
10
l
i
)
s
h
g
e
y
a
r
(
s
s
e
n
t
h
g
i
r
B
1
RMars
Planetocentric Distance (x103 km)
10
Figure 5: H I Lyman-b was detected along the full length of the Alice slit for both offsets. The
data are shown as a histogram: black, 2.5◦ offset, red: 7.5◦ offset. The errors shown are statistical
in the count rate. Optically thin Chamberlain models without satellite orbits for T=200 K (solid
line) and 260 K (dashed line) are shown, superimposed on a 1 rayleigh interplanetary background.
The radius of Mars is indicated.
22
l
i
)
s
h
g
e
y
a
r
(
s
s
e
n
t
h
g
i
r
B
Lyman−a
10000
1000
100
RMars
Planetocentric Distance (x103 km)
10
Figure 6: Same as Fig. 5 for H I Lyman-a
. The shape of this profile is very similar to the slant
intensity profiles derived from the Mariner 6 and 7 fly-bys by Barth et al. (1971), although lower
in absolute brightness as the Mariner fly-bys occurred at a time of high solar activity. The same
models are shown, superimposed on an interplanetary background of 250 rayleighs.
23
)
1
−
Å
s
h
g
e
y
a
R
l
i
(
s
s
e
n
t
h
g
i
r
B
)
1
−
Å
s
h
g
e
y
a
R
l
i
(
s
s
e
n
h
g
i
r
t
B
Lyg
Lyb
OI
row 15
1000
1200
Wavelength (Å)
1400
1600
Lyg
Lyb
OI
row 16
5
4
3
2
1
0
5
4
3
2
1
0
1000
1200
Wavelength (Å)
1400
1600
Figure 7: Extracted limb scan spectra. For row 15 (top), the photon counts from the first 12
exposures (376 s, see Fig. 3) were co-added. For row 16 (bottom), the last 14 exposures (476 s)
were co-added. Only H I and O I emissions are detected.
24
)
m
k
(
e
d
u
t
i
t
l
A
1400
1200
1000
800
600
400
200
1
10
OI 1304 Brightness (Rayleighs)
100
1000
Figure 8: Extracted O I l 1304 brightness from rows 14 -- 17 as a function of altitude above the
Martian limb. The vertical bars illustrate the extent of the trapezoidal altitude weighting function
for each row while the horizontal bars are the statistical uncertainty in the count rate. The two-
component oxygen model shown is described in the text.
25
|
1502.05035 | 1 | 1502 | 2015-02-17T20:59:18 | The APOGEE Spectroscopic Survey of Kepler Planet Hosts: Feasibility, Efficiency, and First Results | [
"astro-ph.EP",
"astro-ph.SR"
] | The Kepler mission has yielded a large number of planet candidates from among the Kepler Objects of Interest (KOIs), but spectroscopic follow-up of these relatively faint stars is a serious bottleneck in confirming and characterizing these systems. We present motivation and survey design for an ongoing project with the SDSS-III multiplexed APOGEE near-infrared spectrograph to monitor hundreds of KOI host stars. We report some of our first results using representative targets from our sample, which include current planet candidates that we find to be false positives, as well as candidates listed as false positives that we do not find to be spectroscopic binaries. With this survey, KOI hosts are observed over ~20 epochs at a radial velocity precision of 100-200 m/s. These observations can easily identify a majority of false positives caused by physically-associated stellar or substellar binaries, and in many cases, fully characterize their orbits. We demonstrate that APOGEE is capable of achieving RV precision at the 100-200 m/s level over long time baselines, and that APOGEE's multiplexing capability makes it substantially more efficient at identifying false positives due to binaries than other single-object spectrographs working to confirm KOIs as planets. These APOGEE RVs enable ancillary science projects, such as studies of fundamental stellar astrophysics or intrinsically rare substellar companions. The coadded APOGEE spectra can be used to derive stellar properties (T_eff, log(g)) and chemical abundances of over a dozen elements to probe correlations of planet properties with individual elemental abundances. | astro-ph.EP | astro-ph |
Accepted in AJ - 16 Feb. 2015
Preprint typeset using LATEX style emulateapj v. 5/2/11
THE APOGEE SPECTROSCOPIC SURVEY OF KEPLER PLANET HOSTS: FEASIBILITY, EFFICIENCY,
AND FIRST RESULTS
Scott W. Fleming1,2, Suvrath Mahadevan3,4, Rohit Deshpande3,4, Chad F. Bender3,4, Ryan C. Terrien3,4, Robert
C. Marchwinski3,4, Ji Wang5, Arpita Roy3,4, Keivan G. Stassun6,7, Carlos Allende Prieto8,9, Katia Cunha10,11,
Verne V. Smith12, Eric Agol13, Hasan Ak14,3, Fabienne A. Bastien3,15, Dmitry Bizyaev16, Justin R. Crepp17, Eric
B. Ford3,4, Peter M. Frinchaboy18, Domingo An´ıbal Garc´ıa-Hern´andez8,9, Ana Elia Garc´ıa P´erez19, B. Scott
Gaudi20, Jian Ge21, Fred Hearty3, Bo Ma21, Steve R. Majewski19, Szabolcs M´esz´aros22, David L. Nidever23,19,
Kaike Pan16, Joshua Pepper24,6, Marc H. Pinsonneault20, Ricardo P. Schiavon25, Donald P. Schneider3,4, John
C. Wilson19, Olga Zamora8,9, Gail Zasowski26,27
Accepted in AJ - 16 Feb. 2015
ABSTRACT
The Kepler mission has yielded a large number of planet candidates from among the Kepler Objects
of Interest (KOIs), but spectroscopic follow-up of these relatively faint stars is a serious bottleneck in
confirming and characterizing these systems. We present motivation and survey design for an ongoing
project with the SDSS-III multiplexed APOGEE near-infrared spectrograph to monitor hundreds of
KOI host stars. We report some of our first results using representative targets from our sample, which
include current planet candidates that we find to be false positives, as well as candidates listed as false
positives that we do not find to be spectroscopic binaries. With this survey, KOI hosts are observed
over ∼ 20 epochs at a radial velocity precision of 100 − 200 m s−1. These observations can easily
identify a majority of false positives caused by physically-associated stellar or substellar binaries, and
in many cases, fully characterize their orbits. We demonstrate that APOGEE is capable of achieving
RV precision at the 100 − 200 m s−1 level over long time baselines, and that APOGEE's multiplexing
capability makes it substantially more efficient at identifying false positives due to binaries than
other single-object spectrographs working to confirm KOIs as planets. These APOGEE RVs enable
ancillary science projects, such as studies of fundamental stellar astrophysics or intrinsically rare
substellar companions. The coadded APOGEE spectra can be used to derive stellar properties (Teff ,
log g) and chemical abundances of over a dozen elements to probe correlations of planet properties
with individual elemental abundances.
[email protected]
1 Computer Science Corporation, 3700 San Martin Dr, Balti-
more, MD, 21218, USA
2 Space Telescope Science Institute, 3700 San Martin Dr, Bal-
timore, MD, 21218, USA
3 Department of Astronomy and Astrophysics, The Pennsyl-
vania State University, 525 Davey Laboratory, University Park,
PA 16802, USA
4 Center for Exoplanets and Habitable Worlds, The Pennsyl-
vania State University, University Park, PA 16802, USA
5 Department of Astronomy, Yale University, New Haven, CT
06511, USA
6 Department of Physics & Astronomy, Vanderbilt University,
Nashville, TN 37235, USA
7 Department of Physics, Fisk University, Nashville, TN
37208, USA
8 Instituto de Astrof´ısica de Canarias (IAC), E-38200 La La-
guna, Tenerife, Spain
9 Departamento de Astrof´ısica, Universidad de La Laguna, E-
38206 la Laguna, Tenerife, Spain
10 Observatorio Nacional-MCTI, Rio de Janeiro, RJ 20921-
400, Brazil
11 Steward Observatory, University of Arizona, Tucson, AZ
85721, USA
12 National Optical Astronomy Observatory, 950 N. Cherry
Avenue, Tucson, AZ, 85719, USA
13 Department of Astronomy, Box 351580, University of Wash-
ington, Seattle, WA 98195, USA
14 Faculty of Sciences, Department of Astronomy and Space
Sciences, Erciyes University, 38039 Kayseri, Turkey
15 Hubble Fellow
16 Apache Point Observatory, P.O. Box 59, Sunspot, NM
88349-0059, USA
17 University of Notre Dame, Department of Physics, 225
Nieuwland Science Hall, Notre Dame, IN 46556, USA
18 Department of Physics and Astronomy, Texas Christian
University, Fort Worth, TX 76129, USA
19 Department of Astronomy, University of Virginia, Char-
lottesville, VA 22904-4325, USA
20 Department of Astronomy, The Ohio State University, 140
West 18th Avenue, Columbus, OH 43210, USA
21 Astronomy Department, University of Florida, 211 Bryant
Space Sciences Center, Gainesville, FL 32111, USA
22 ELTE Gothard Astrophysical Observatory, H-9704 Szom-
bathely, Szent Imre herceg st. 112, Hungary
23 Department of Astronomy, University of Michigan, 830
Dennison, 500 Church St., Ann Arbor, MI 48109-1042, USA
24 Lehigh University, Department of Physics, 16 Memorial
Drive East, Bethlehem, PA 18015, USA
25 Astrophysics Research Institute, IC2, Liverpool Science
Park, Liverpool John Moores University, 146 Brownlow Hill, Liv-
erpool, L3 5RF, UK
26 Department of Physics & Astronomy, Johns Hopkins Uni-
versity, Baltimore, MD 21218, USA
27 NSF Astronomy and Astrophysics Postdoctoral Fellow
2
Scott W. Fleming et al.
1. INTRODUCTION
1.1. Kepler's Planet Candidates
2011).
2010; Gilliland et al.
The Kepler spacecraft's primary mission is to deter-
mine the frequency of Earth-sized exoplanets orbiting in
the habitable zone of their parent stars (Borucki et al.
2010; Koch et al.
2010), with a second objective of
studying a wide variety of stellar astrophysics via as-
teroseismology (e.g., Chaplin et al.
In ad-
dition,
the high precision photometry (∼ 80 ppm
over 6-hour timescales for the brightest (Kp . 15)
dwarfs, Caldwell et al.
2011;
Christiansen et al. 2012) enables studies of giant exo-
planets and a wide variety of variable stars. Its photo-
metric band Kp covers 423 - 897 nm and is similar to, but
broader than, a combined V and R band (Koch et al.
2010). To find exoplanets, Kepler makes use of the tran-
sit method, which detects planet candidates by mea-
suring the flux loss that occurs when a planet crosses
the face of its parent star. However, there are several
sources of false positives that must be taken into account
when analyzing these candidates, most notably: graz-
ing eclipsing binaries (EBs), EBs (including hierarchical
triples) whose eclipse depths are diluted by another star
through flux contamination, brown dwarfs or low mass
stars that have radii comparable to giant exoplanets, and
even larger exoplanets that transit a fainter star within
the photometric aperture.
Because of these sources of false positives, the Kepler
team makes a very clear distinction between candidate
exoplanets and those that have been dynamically con-
firmed through spectroscopic radial velocity (RV) mea-
surements or through photodynamical modeling (e.g.,
Holman et al. 2010; Carter et al. 2011). Kepler Ob-
jects of Interest (KOIs) consist of candidate exoplanets,
eclipsing binaries, and known false positives. Those KOIs
that are not known to be false positives or EBs are re-
ferred to as "active planet candidates" (Borucki et al.
2011a,b; Batalha et al.
2013), but for simplicity, we
will refer to such Kepler planet candidates as "KPCs"
throughout the rest of this paper. An intermediate level
of classification consists of "validated" exoplanets, which
have very low probabilities of being blended EBs as de-
termined through a Monte Carlo statistical analysis of
the Kepler photometry (e.g., Torres et al. 2011).
As of October 2014, there are a total of 4229 KPCs
amongst 3251 Kepler stars 1 , but only ∼ 20% (653) of
the stars host multiple KPCs.
It is estimated that as
many as 15-26% of transiting planets may have clearly
detected transit timing variations (Ford et al. 2012),
which allow for mass determinations photometrically.
Even still, a majority of KPCs will require RV obser-
vations to confirm their planetary nature. Such time-
series RV observations are resource intensive, so effi-
cient identification of false positive candidates is neces-
sary to ensure efficient follow-up of likely planets.
In
addition to aiding in the confirmation of KPCs, robustly
determining the false positive rate amongst KPCs is re-
quired when conducting statistical analyses of this pop-
ulation. A number of studies have attempted to per-
form such analyses,
including investigations of planet
1
http://exoplanetarchive.ipac.caltech.edu/cgi-bin/
ExoTables/nph-exotbls?dataset=cumulative_only
frequency as functions of orbital periods and stellar
host properties (Borucki et al.
2011b; Youdin 2011;
Howard et al. 2012), and studies of the eccentricity dis-
tribution (Moorhead et al. 2011).
Aside from the false positive rate of KPCs, knowledge
of the host star(s) intrinsic properties (e.g., mass, radius,
effective temperature, surface gravity, metallicity) is re-
quired to determine the masses and radii of the exoplan-
ets, as well as to conduct studies of planetary properties
as functions of these stellar parameters. The Kepler In-
put Catalog (KIC, Brown et al. 2011) provides a pho-
tometrically derived Teff , log g, [Fe/H] and E(B−V) for
every star within Kepler 's field of view through a combi-
nation of calibrated fluxes using {g, r, i, z} filters similar
to the original SDSS filters (Fukugita et al. 1996) and a
narrow-band D 51 filter modeled after the Dunlap Obser-
vatory DD 51 filter. The catalog was originally used to
inform target selection for the mission, but in the absence
of a comprehensive spectroscopic survey of all ∼ 150, 000
Kepler stars, the catalog's stellar parameters have been
used in analyses of planet candidates. There are ongoing
efforts to provide improved stellar parameters of Kepler
targets by aggregating photometry, spectroscopy, aster-
oseismology, and transit analyses (Huber et al. 2014).
The majority of false positive KPCs are expected to be
caused by astrophysical sources rather than random or
systematic errors, specifically, EBs whose eclipse depths
are similar to that expected from a transiting planet
(Borucki et al. 2011b). Fig. 1 demonstrates six of the
most common sources of transiting KPC scenarios. In
each panel, the larger (yellow) star is the suspected KPC
host, and all objects within the panels are assumed to be
within the aperture used to create the Kepler lightcurve.
Each Kepler "optimal aperture" is variable, but is typ-
ically many arcseconds in size (Twicken et al. 2010).
The dashed circles represent a spectrograph fiber's field-
of-view (FoV, not to scale). The titles in each panel also
denote, qualitatively, how often the given scenarios can
be characterized by time-series RVs at modest precision
(∼ 100 m s−1 level). Note that in addition to stellar
eclipses being diluted to look like giant planets, transits
of giant planets can also be diluted to look like smaller
planets.
1.2. Sources Of False Positives
Scenarios 1, 2, 4 and 6 all involve a physical companion
orbiting the KPC host star. In these scenarios, RV obser-
vations can detect the presence of grazing EBs (Scenario
1), EBs that are diluted by light from a third star within
the Kepler aperture, but resolved on-sky with the spec-
trograph (Scenario 2), or consist of a very low-mass star
(VLMS) or brown dwarf companion (Scenario 4), a ma-
jority of the time. An important sub-category of Scenar-
ios 1 and 4 include EBs whose orbits produce only a sec-
ondary eclipse and no primary eclipse (Santerne et al.
2013). These false positives may be more common for
longer period KPCs, where a companion star in an ec-
centric orbit is more likely to undergo secondary eclipse
near periastron, but exhibit no primary eclipse. In addi-
tion, the most massive, bona fide planets at short orbital
periods will induce a Doppler velocity shift detectable at
the ∼ 100 m s−1 level (Scenario 6). In those rare cases,
their planetary nature will be confirmed through the
APOGEE RVs by phasing them to the Kepler -derived
APOGEE Spectroscopy of the Kepler Field
3
!"#Grazing EB - Most
2. Diluted EB - Most
3. Unresolved EB - None
4. VLMS / BD - Most
5. Close Triples - Some
6. Planet - Rarely
Figure 1. The most common scenarios that can produce a
lightcurve consistent with a transiting planet. The titles qualita-
tively identify those scenarios that can be detected by a RV survey
at the level of ∼ 100 m s−1 ("Most", "Some", etc.). In each panel,
the larger (yellow) star is the assumed KPC host star, while the
dashed circles represent a spectrograph's FoV (not to scale). All
sources in each panel lie within the Kepler photometric aperture,
which is typically many arcseconds in size. In Scenario 3, the term
"unresolved" refers to the fact that the EB is unresolved in the
Kepler aperture. In Scenario 6, only short-period, massive planets
would be detected with APOGEE. Note that giant planets can also
be diluted to look like smaller-sized planets in these scenarios.
orbital period.
Stellar systems that are either physical multiples or vi-
sual companions with small separations on-sky, and con-
sist of an EB, are represented by Scenario 5.
In this
scenario, the diluted EB is only detectable if the flux of
at least one component of the EB pair is sufficiently high
that it appears in the cross-correlation function, or if the
combined mass of the EB pair induces a sufficient velocity
shift on the third (brightest) star in the system. When
the EB pair is composed of cooler K or M dwarf stars in
the presence of a hotter primary, they are easier to detect
in the NIR than in the optical, since the flux contrast is
reduced in the H band (e.g., Kepler-16, Bender et al.
2012). For a binary system composed of dwarf stars at a
signal-to-noise of 100, secondaries with mass ratios down
to ∼ 0.1 are detectable in the H band (Bender & Simon
2008), while mass ratios are limited to ∼ 0.5 in the op-
tical. The detection limit for a given system depends
on the number of stellar components within the aperture
(e.g., is it a binary versus a triple system?), and whether
any of those components are evolved (observing in the
NIR is beneficial for components of differing Teff ratios,
not for brightness differences due to differing radii).
Scenario 3 represents a Kepler -unresolved EB, where
the variable star is within the Kepler aperture, but is
exterior to the spectrograph's FoV relative to the KPC
host star. This is the only scenario where RV observa-
tions will be not be able to detect any false positives,
unless the RV survey targets every star within a given
KPC's Kepler aperture. Fortunately, Scenarios 2, 3 and
5 can sometimes be tested photometrically with time-
series photometry from the ground at greater spatial res-
olutions (e.g., Col´on et al. 2012). In addition, these are
also the scenarios that are more likely to be solved us-
ing Kepler data alone, e.g., by searching for flux centroid
shifts.
1.3. Paper Outline
2010)
In this paper, we introduce our program to ob-
serve hundreds of KPCs using the Apache Point Ob-
servatory Galactic Evolution Experiment (APOGEE,
Majewski et al.
spectrograph (Wilson et al.
2010, 2012) on the Sloan 2.5 m telescope (Gunn et al.
2006), recently finished as part of the Sloan Digital Sky
Survey III (SDSS-III, Eisenstein et al. 2011) and con-
tinuing in SDSS-IV (2014-2020). Our program provides
an efficient means of determining the false positive rate
of KPCs due to physically-associated binary stellar sys-
tems. At the same time, these spectra are used in a
variety of projects concerning the false positives them-
selves, including characterization of the orbits and mea-
surements of the mass ratios for many of the spectro-
scopic binaries (SBs), or orbital characterization of in-
trinsically rare, massive (M & 10 MJup), substellar com-
panions such as brown dwarfs and massive gas giant plan-
ets (Marcy & Butler 2000; Sahlmann et al. 2011). For
KPCs that remain viable, host star properties such as
Teff, log g, and chemical abundances for dozens of ele-
ments can be derived using the APOGEE spectra.
In Section 2 we describe the APOGEE instrument and
main survey, the methods used to derive RVs from its
spectra, and its current RV precision floor. In Section
3.1 we present RVs of five current and former KOIs ob-
served during SDSS-III. Three of these happened to be
observed as part of a separate APOGEE EB program,
and we present some conclusions on the nature of those
KOIs as a precursor to our larger KPC campaign. We
also present the first results from our APOGEE-Kepler
KOI campaign, using the (since confirmed) exoplanet
host KIC 6448890 to test our long-term RV precision,
and definitively identifying KIC 6867766 as a false posi-
tive exoplanet.
In Section 3.2 we compare the efficiency of a sur-
vey using a high resolution, NIR, multi-object spec-
trograph against other planet-hunting spectrographs:
HARPS-North, which is a clone of HARPS-South with
some improvements (Mayor et al. 2003), Keck HIRES
(Vogt et al. 1994), SOPHIE (Perruchot et al. 2008),
and HET HRS (Tull 1998). We demonstrate that by
using a multiplexing instrument in the NIR to conduct
a survey at modest RV precision (100 m s−1), false pos-
itives can be identified more efficiently compared to the
single-object instruments, reserving telescope time on
those other resources for confirmation of the remaining
KPCs at significantly higher precision. In Section 4 we
review other techniques for determining the false positive
rate of Kepler KPCs, and highlight the science enabled
by extracting abundances from the coadded APOGEE
spectra. We summarize our findings in Section 5.
2. APOGEE SURVEY OVERVIEW
APOGEE is a survey of Milky Way stars using a multi-
object, fiber-fed, NIR spectrograph housed in a vacuum
cryostat, that can observe up to 300 objects simultane-
ously, producing R ∼ 22500 spectra covering a wave-
length range of 1.51−1.68 µm using a volume phase holo-
graphic grating mosaic. Details of the instrument design
can be found in Wilson et al. (2010) and Wilson et al.
(2012). Typically the instrument achieves a signal-to-
noise ratio per pixel of 100 (∆λ ∼ 0.1 − 0.17A) on an
H = 11 star in a single visit (one hour of total in-
tegration). Most stars are observed on a minimum of
4
Scott W. Fleming et al.
ing. Critically, these pipeline-derived RVs only work for
simple cross-correlation functions, and are expected to
fail when there is contamination from multiple stellar
spectra, as in the case for most binary stars. As such, we
derive our own RVs using additional, interactive process-
ing of the data. These steps include manually correcting
residual OH sky emission lines, selecting templates from
a finer grid, and interactively fitting cross-correlation
peaks, which may often be asymmetric or have multi-
ple components in the case of binary stars. We cal-
culate uncertainties for our RV measurements following
the maximum-likelihood procedure laid out by Zucker
(2003). This approach derives an analytical relationship
between the cross-correlation function and it's first and
second derivatives to account for uncertainty contribu-
tions related to the sampling and sharpness of the corre-
lation peak, and the signal-to-noise of the target and tem-
plate spectra. The RVs are then fit using a custom wrap-
per to the RVLIN software package (Wright & Howard
2009), which includes the ability to fit both components
of a double-line spectroscopic binary through an itera-
tive approach, and forces some orbital parameters to be
identical between both components (e.g., orbital period,
eccentricity, epoch of periastron). In some cases we make
use of our IDL-based Levenberg-Marquardt fitting code
used in Bender et al. (2012).
3. RESULTS
3.1. Initial Case Studies
We have selected five KOIs with diverse histories
and current statuses to test and develop our analysis
pipelines. Three of these targets were observed as part
of an SDSS-III ancillary program studying Kepler EBs
(Mahadevan et al. 2015) (hereafter MAH2015). These
targets were at one point Kepler planet candidates,
but were determined to be likely EBs by the time the
MAH2015 observations began. Note that since these
KOI hosts were observed through a different program,
the total number of epochs for these targets is less than
the number of epochs that the SDSS-Kepler KOI pro-
gram obtains (the SDSS-III EB program obtained 3 − 6
epochs for each target, compared to > 18 epochs for
the SDSS-III KOI program). The two other KPC hosts
presented here come from our SDSS-III Kepler KOI pro-
gram. We summarize our findings on these targets to
demonstrate the diversity of astrophysical configurations
encountered in our spectroscopic observations.
3.1.1. SDSS-III EB Program: KIC 1571511
KIC 01571511 (KOI 362, Kp = 13.42, H =
12.04), consists of an F-type dwarf and low-mass M
dwarf (Ofir et al. 2012), and was observed as part of
MAH2015. The orbital period is 14.0224519 days and
the radius as estimated in the NExScI KOI catalog2 is
14.7 ± 6.4 R⊕. The eclipses of such a low-mass star
(∼ 2% decrease in flux during primary eclipse) are com-
parable to those expected for a gas giant planet. Indeed,
this star was originally suspected to be an overlooked
gas giant exoplanet (Coughlin et al. 2011; Ofir et al.
2012). In the specific case of KIC 01571511, there is a
small secondary eclipse (∼ 0.05%) detected in the Kepler
2 http://exoplanetarchive.ipac.caltech.edu
Figure 2. Kepler module footprints with APOGEE FoV (circles,
2.98◦ diameters) from our SDSS-III Kepler EB and KOI programs.
Five additional Kepler modules will be observed by APOGEE in
SDSS-IV.
three different nights, so that short-period binaries can
be flagged. Each field on the sky is normally observed
in multiples of three, ranging from 3 to 24 epochs, with
brighter targets swapped for new stars after three obser-
vations. Aluminum plug plates hold optical fibers that
carry the star light from the telescope into the instru-
ment. The primary science goal of the survey is to study
the Milky Way by measuring radial velocities and chem-
ical abundances of ∼ 105 red giant stars, but a vari-
ety of additional science projects are included. A sum-
mary of the project can be found in Allende Prieto et al.
(2008b). A detailed description of the survey will ap-
pear in Majewski et al.
(2015). Details of the target
selection for the survey in SDSS-III can be found in
Zasowski et al. (2013).
The telescope's field-of-view (FoV) covers a circular
area 1.49◦ in radius, which matches well to the size of
a given Kepler module. Fig. 2 shows the Kepler mod-
ules' footprints, along with the three SDSS FoVs for
our programs observed during SDSS-III. A total of 163
KPC host stars are observed in the SDSS-III KOI field
(blue). Those targets were selected from all KPCs that
had H < 14, 153 of which were dispositioned as planet
"candidates" as of August 2013 (four others were con-
firmed exoplanets, six were not dispositioned yet). As
can be seen, a single SDSS footprint covers most of a
Kepler module's FoV. In addition to the KPC hosts ob-
served during SDSS-III, five additional Kepler modules
will be observed in SDSS-IV.
The APOGEE data processing pipeline is described
in Nidever et al.
(2015). Basic steps include collaps-
ing the detector exposures for each of the three NIR
arrays from 3D data cubes to 2D images, flat fielding,
aperture extraction, wavelength calibration, sky subtrac-
tion, telluric correction, and measurement of RVs. The
RVs currently calculated by the automated pipeline make
use of a grid of synthetic spectra calculated from AT-
LAS9 stellar model atmospheres (M´esz´aros et al. 2012;
Zamora et al. 2015). The mean, internal RV precision
of the pipeline-produced relative RVs is ∼ 100 m s−1, al-
though it does depend on signal-to-noise, spectral type,
and level of residual systematics from the data process-
APOGEE Spectroscopy of the Kepler Field
5
Figure 3. APOGEE RVs of the Kepler EB KIC 01571511 and the
best-fit model from Ofir et al.
(2012) shown as the solid line. A
constant offset between the model and the APOGEE RVs has been
applied to account for a zero-point offset. For the given Kepler pe-
riod and epoch of transit, it is clear even with just three APOGEE
RVs that the object is not a planet, because the change in RV over
a short fraction of the orbit is much greater (∼ 10 km s−1) than
expected for a planetary mass.
lightcurve, which can be used to derive an estimate of the
relative Teff ratio between the primary and secondary,
and can therefore be used to help determine whether the
object is a likely stellar companion. However, there is
no guarantee that an EB system with a primary eclipse
will also show a secondary eclipse, nor that the secondary
eclipse is detectable even with Kepler's precision. In fact,
Santerne et al. (2012) found that some of their false pos-
itive KPCs were EBs in eccentric orbits for which only
the shallower, secondary eclipses are present, but were
mistaken as planetary transits across the primary.
Fortunately, these EBs are fairly trivial to detect spec-
troscopically, as demonstrated in Fig. 3.
In this fig-
ure, we plot the best-fit RV model from the analysis
by Ofir et al.
(2012), noting that there is a typo in
the value of ω1 in their Table 3 that is missing a mi-
nus sign. We also plot the three APOGEE RVs obtained
through the ancillary program (Table 1). Only a con-
stant offset between the model and APOGEE data is
included to account for instrumental zero-point differ-
ences. Even with three data points, the RV variation
observed in the APOGEE RVs is inconsistent with a gi-
ant planet, given the period and epoch of transit from
the Kepler lightcurve, because the change in RV over a
short fraction of the orbit is much greater (∼ 10 km s−1)
than expected for a planetary mass. These data also
demonstrate that APOGEE is capable of producing RVs
at the ∼ 100 − 200 m s−1 for H ∼ 12 stars based on the
rms residual to the well-determined orbital solution from
Ofir et al. (2012). KIC 1571511 corresponds to Scenario
4 in Fig. 1, where a low-mass star generates an eclipse
depth comparable to that expected from a giant planet.
3.1.2. SDSS-III EB Program: KIC 3848972
KIC 3848972 (KOI 1187, Kp = 14.49, H = 12.80) is
listed in both the EB and KOI catalogs, and was observed
as part of MAH2015. As a KOI, the target was listed as
a "False Positive" in the Q1-Q8 catalog, but is currently
absent in the Q1-Q16 catalog. The KOI Q1-Q8 catalog
Figure 4. Orbital phase-folded RVs for KIC 3848972, using the
period and ephemeris values from the KOI Q1-Q8 Catalog (left)
and EB Catalog (right). The period ambiguity arises from whether
the system is treated as an object that only produces a primary
transit (KOI solution), or an eclipsing binary that produces both a
primary and secondary eclipse with similar depths (EB solution).
No significant RV variation is seen beyond a few hundred m s−1,
disfavoring a physically bound stellar companion at either of these
orbital periods as the source of the Kepler signal.
lists a period of P = 0.37052915 days, while the EB cat-
alog lists a period that is twice as long (P = 0.741057
days). The estimated radius reported in the KOI cata-
log is 3.53 ± 0.93 R⊕. Where the EB Catalog assumes
two nearly-equal eclipses from primary and secondary
eclipse events, the KOI catalog reports half the orbital
period and defines the secondary eclipse as undetected
or absent. Multi-color, ground-based photometry was
observed by Col´on et al.
(2012) using a tunable filter
on the OSIRIS instrument on the 10.4-m Gran Telesco-
pio Canarias (GTC). They find a consistent star-planet
radius ratio (2σ) in both their blue and red filters, but
measure a statistically significant (5.8σ) difference in the
eclipse depths. Interestingly, the color differences during
eclipse suggest that the secondary component is bluer
than the primary.
Only three APOGEE spectra were obtained for this
target as part of MAH2015, however, a check on bina-
rity can still be performed provided the orbital phase
coverage is reasonable. We conduct a one-dimensional
cross-correlation using a K-type dwarf template (Teff =
5000K). We do not see evidence for any significant ro-
tational broadening greater than ∼ 10 − 20 km s−1. Al-
though the APOGEE spectra for this star are somewhat
noisy, we find a single, very stable CCF peak with no
RV variation greater than a few hundred m s−1 (Table
1). There is no obvious correlation with orbital phase
after folding on both the KOI and EB Catalog periods
and ephemerides, despite spanning ∼ 80% of the KOI or-
bital phase and ∼ 20% of the EB Catalog orbital phase,
respectively (Fig. 4). If the signal was caused by a hot-
ter (bluer) secondary orbiting a brighter primary, the
expected RV amplitude should be many tens of km s−1.
Another possible explanation is that the observed Ke-
pler signal comes from an object transiting a low-mass,
cool (red) star that is within the photometric aperture
(constrained to be ra < 2′′ given the Col´on et al. (2012)
GTC aperture), but too faint to detect spectroscopically
with APOGEE in the presence of the bright primary
star.
In this case, the color-dependent transit depths
are caused because the fainter, redder component is the
one being transited, hence the overall color of the com-
bined light appears to shift towards the blue during the
transit event. Intrigued by this possibility, we obtained
Keck adaptive optics (AO) imaging to search for a fainter
companion that might be the source of the Kepler signal.
The AO image was acquired on UT 2014 Jul 17,
6
Scott W. Fleming et al.
Morton & Johnson (2011), we can estimate the prob-
ability of having a blend source within a given aper-
ture using a model of the galactic population from TRI-
LEGAL3 (Girardi et al. 2012). We generate a TRI-
LEGAL population in a one square degree area cen-
tered on KIC 3848972 with the default settings (see
Appendix A). We calculate a mean stellar density
of 0.004581 stars arcsec−2 within Kepler magnitudes
15.49 ≤ Kp ≤ 21.26. This magnitude range is cho-
sen because it represents stars faint enough to be un-
detected in the Kepler aperture but still able to pro-
duce a transit depth of δ = 0.00196. With this mean
density, the probability of just finding a potential blend
source within this magnitude range is 5.8% within 2′′(the
GTC photometric aperture) and just 0.36% within 0.′′5
(the area least probed by the Keck AO images, Fig. 5).
From Morton & Johnson (2011), the probability that
this blend source is an EB with a configuration that
could mimic a transiting planet signal is on the order
of 2.5E-4, so the probability of having a background EB
as the source of this KPC is on the order of {9E-7, 1.4E-
5} within {0.5, 2.0} arcseconds, respectively. Thus, the
more likely EB false positive scenario is a bound eclipsing
binary causing the transit signal.
The lack of observed RV variability indicates this KOI
is not due to a physically bound stellar companion orbit-
ing the brightest component of the KIC 3848972 system,
while the Keck AO images constrain any bound, diluted
EB to be either within ∼ 0.5′′ of the primary, or more
than 5 magnitudes fainter than the primary in the Ks
band. Given the Col´on et al.
(2012) observations, we
hypothesized that this KOI corresponded to Scenario 2
or 5 in Fig. 1, but our observations rule out Scenario
2 and tightly constrain the separation of a diluted EB
under Scenario 5.
3.1.3. SDSS-III EB Program: KIC 3861595
KIC 3861595 (KOI 4, Kp = 11.43, H = 10.27) is listed
in both the EB and KOI catalogs, and was observed as
part of MAH2015. As a KOI, the target was initially
listed as a "False Positive" in the Q1-Q8 catalog, and
is currently listed as "Not Dispositioned" in the Q1-Q16
catalog. The orbital period is 3.8493724 days and the
estimated planet radius is 11.8 ± 1.6 R⊕. Some ground-
based observations have been conducted and reported
at the Kepler Community Follow-up Program (CFOP)
website4. These include several optical spectra from the
TRES spectrograph (Szentgyorgyi & Fur´esz 2007) that
indicated the star was a rapid rotator (40-50 km s−1),
and potentially variable at a level of a few hundreds of
m s−1.
Imaging from the 1-m Nickel telescope at Lick
Observatory and Keck HIRES guider images show two
nearby stars within ten arcseconds of the target. Both
nearby stars appear to be approximately 6 magnitudes
fainter than the target.
In addition to six APOGEE spectra, MAH2015 ob-
tained five optical spectra for this target using the High
Resolution Spectrograph (HRS) on the Hobby-Eberly
Telescope. The HRS was used in the 30,000 resolution
mode, with the 316 g/mm grating at a central wave-
3 http://stev.oapd.inaf.it/cgi-bin/trilegal_1.6
4 https://cfop.ipac.caltech.edu/home/
Figure 5. Upper limits on secondary companion mass from Keck
AO imaging. The photometric aperture used by Col´on et al.
(2012) is marked by the vertical line, and serves as the outer limit
of where the eclipsing object might lie.
using NIRC2 (instrument PI: Keith Matthews) and
the Keck II Natural Guide Star (NGS) AO system
(Wizinowich et al. 2000). We used the narrow cam-
era setting with a plate scale of 10 mas pixel−1, which
provides a fine spatial sampling of the instrument point
spread function (PSF). The observing conditions were
excellent, with a seeing of 0′′.3. KIC 3848972 was ob-
served at an airmass of 1.12. We used the KS filter to
acquire the image using a 3-point dither method. At each
dither position, we took a total of 10 coadds composed
of 5-second exposures. The total on-source integration
time is therefore 150 seconds.
The raw NIRC2 data were processed using standard
techniques to replace bad pixels, flat-field, subtract ther-
mal background, align, and coadd frames. We calculated
the 5σ detection limit as follows. We first defined a series
of concentric annuli centered on the star. For the con-
centric annuli, we calculated the median and standard
deviation of flux for pixels within these annuli. We define
the 5σ detection limit as five times the standard devia-
tion above the median flux. Representative 5σ detection
limits are {1.6, 3.3, 5.0, 5.4} magnitudes for projected
separations of {0′′.1, 0′′.2, 0′′.5, 1′′}, respectively.
We translate the 5σ upper limits on companion bright-
ness into upper limits on companion mass using the SED
models compiled in Kraus & Hillenbrand (2007), assum-
ing the secondary is a bound companion and that differ-
ential extinction between the two spectral types is mini-
mal in the Ks band. For a given absolute Ks magnitude
of the primary, the contrast curve from the Keck AO data
gives a lower limit for the secondary's absolute Ks mag-
nitude, which we then interpolate into a mass using the
Kraus & Hillenbrand (2007) models. We adopt primary
spectral types of G0 and K5 as conservative upper and
lower limits based on our spectroscopic cross-correlation
analysis. We are able to rule out any bound companions
more massive than 0.2 M⊙ at the 5σ level exterior to 0.2
arcseconds (Fig. 5).
the transit signal could be caused
by a fainter background or foreground star that is
physically unassociated with KIC 3848972, but still
within the Kepler photometric aperture.
Following
Alternatively,
APOGEE Spectroscopy of the Kepler Field
7
length of λ0 = 5936 A. The HET spectra were re-
duced using our optimal extraction pipeline described
in MAH2015. We find a good template match (for both
HET and APOGEE) using a mid-F spectral template
rotationally broadened to 40 km s−1, in agreement with
the CFOP notes from the TRES observations. The es-
timated spectroscopic rotation rate of ∼ 40 km s−1,
combined with an estimated rotation rate of 5.65-5.8
days (Hirano et al. 2012; Rhodes & Budding 2014), re-
sults in an equatorial radius of ∼ 4.5 R⊙, some-
what larger than the spectroscopically determined ra-
dius of 2.727 ± 0.504 R⊙ (Buchhave et al. 2012) and
the Stellar Parameter Catalog's value of 2.992+0.469
−0.743 R⊙
(Huber et al. 2014), which also uses the spectroscopic
stellar parameters of Buchhave et al. (2012), but tie the
stellar parameters to Dartmouth stellar evolution models
(Dotter et al. 2008).
We cross-correlate the HET and APOGEE spectra
with mid-F spectral templates rotationally broadened
to 40 km s−1. For this target, we used subsections of
the APOGEE spectrum (1.515-1.560 µm, 1.586-1.605
µm, 1.615-1.635 µm, and 1.6475-1.6775 µm) to avoid
several of the broadest lines. We find a single-peaked,
broad cross-correlation function, with RV variation at
the ∼ 1 km s−1 level (Table 1). We note that the
RV scatter is larger than most of the A and F stars
observed by Lagrange et al.
(2009). However, phase-
folding and fitting both sets of RVs at the orbital period
and ephemerides found in the KOI and EB catalogs fails
to resolve a signal of orbital motion from a bound com-
panion at that period.
In fact, we find that if we fit each set of RVs separately,
fix the period and ephemeris to the KOI values, and force
the eccentricity to zero, the best-fit RV semiamplitudes
differ by a factor of 2.5 (Fig. 6). A color-dependent
semiamplitude may signal a blended spectrum (Scenar-
ios 2 or 5); the redder component can affect the line
shapes more significantly in the NIR, but such scenarios
are particularly challenging to identify in rapidly rotating
stars using only a handful of observations. Nevertheless,
we undertake a full line bisector analysis of both HET
and APOGEE spectra. After creating custom numerical
stellar template masks for both the HET and APOGEE
wavelength ranges, we calculate the bisectors of the cross-
correlation function, similar to the procedure described
in Wright et al.
(2013). We are limited to using three
of the six APOGEE observations because they were the
only ones observed on the same plug fiber, and there-
fore should have the same intrinsic profile. The bisectors
appear to be varying both in shape and position (indi-
cating a cause other than bulk motion of the primary)
and the bisector inverse slope (BIS) seems well-correlated
with both RV and CCF FWHM (Fig. 7). However, the
rapidly-rotating nature of this star causes difficulty in
establishing the CCF continuum, and complicates bisec-
tor analysis. Any attempt to calculate errors on the BIS
leads to overestimation, and therefore we are hesitant to
quantify this result beyond saying that a blend scenario
is possible.
We can not definitively show the transit signal
is
caused by a spectroscopic blend with so few bisector mea-
surements. Our HET RVs are consistent with a planetary
companion in a circular orbit, but there are not enough
Figure 6. Phase-folded RVs for APOGEE (red) and HET (blue).
Each set were fit with the orbital period and transit ephemeris
fixed to the KOI value. Eccentricity is forced to zero. We find
that while both sets of RVs appear to be in-phase with the or-
bital parameters, the RV semiamplitudes are quite different. This
suggests the spectrum might be blended (Scenarios 2 or 5 in Fig.
1): the APOGEE spectra can be more sensitive to such a blend
if the temperatures of the blended components are different. Line
bisector variations also suggest a blend, but is not definitive due
to the small number of observations. It is also possible that the
uncertainties are underestimated due to the rapid rotation of the
star (40 km s−1).
60
60
50
50
40
40
30
30
20
20
10
10
h
h
t
t
p
p
e
e
D
D
e
e
g
g
a
a
t
t
n
n
e
e
c
c
r
r
e
e
P
P
0
0
−40 −30 −20 −10
−40 −30 −20 −10
Velocity (km/s)
Velocity (km/s)
0
0
80
60
40
20
h
t
p
e
D
e
g
a
t
n
e
c
r
e
P
0
−12−10 −8 −6 −4 −2 0 2
Velocity (km/s)
)
s
/
m
k
(
S
B
I
−15
−20
−25
−30
−35
)
s
/
m
k
(
S
B
I
0
−2
−4
−6
−33 −32 −31 −30
Velocity (km/s)
−33 −32 −31 −30
Velocity (km/s)
)
s
/
m
k
(
M
H
W
F
)
s
/
m
k
(
M
H
W
F
94
92
90
88
86
84
52
50
48
46
−30
Velocity (km/s)
−29
−28
−30
Velocity (km/s)
−29
−28
Figure 7. Bisector analysis for the three APOGEE spectra that
were on a common fiber (top row) and HET spectra (bottom row).
The grey data point in the HET plots represents a low signal-
to-noise observation.
(Left: ) Bisectors of the cross-correlation
function, shifted by measured radial velocities. Colors are based
on bisector inverse slope (BIS) values. Percentage depth is a proxy
for flux, and does not span the full range from 0 to 1 because
of continuum ambiguities.
(Middle: ) Correlation between BIS
and measured RV. (Right: ) Correlation between FWHM of the
CCF, and measured RV. Uncertainties are determined from the
measurement variation between echelle orders (for HET spectra)
and between the three detectors (for APOGEE spectra). Note
that difficulty in defining the CCF continuum probably leads to an
overestimation of the BIS errors.
RVs to make a firm claim. The APOGEE RVs contradict
this claim, but RV uncertainties for rapid rotators have
not been thoroughly vetted, so the APOGEE RV un-
certainties reported in Table 1 could be underestimated.
We also note that an analysis by Rhodes & Budding
(2014) found that an Algol-type background binary, ap-
proximately 6.5 magnitudes fainter but within the Kepler
photometric aperture, could produce a lightcurve similar
8
Scott W. Fleming et al.
φ (Phase-Folded On KOI T0 )
0.20308
0.35796
0.5319
0.68369
KIC 6867766
1.0
0.8
0.6
0.4
0.2
r
e
w
o
P
F
C
C
d
e
z
i
l
a
r
o
N
Figure 8. The long-term RV rms of the confirmed exoplanet host
star KIC 68448890 demonstrates we are able to achieve relative RV
precision at the 100 m s−1 level.
to what's observed. This scenario, which corresponds to
Scenario 3 in Fig. 1, also remains a possibility given the
companions seen in the Lick and Keck images at this
approximate flux ratio. This KOI is a prime example
as to why it is sometimes necessary to obtain multiple
spectra when searching for exoplanet false positives; ob-
taining just a few spectra, even at orbital quadratures,
may not be sufficient to confidently identify a blended
stellar binary, especially for systems that are rapid rota-
tors. Additional spectroscopic observations will be able
to study the line bisectors and RVs in sufficient detail to
determine the nature of this intriguing KOI.
3.1.4. SDSS-III KOI Program: KIC 6448890
KIC 6448890 (KOI 1241, Kp = 12.44, H = 10.33) is
a system with two exoplanets that have been confirmed
via transit timing variations (Steffen et al. 2013). The
two planets have orbital periods of 10.5016 and 21.40239
days, radii of 0.581 and 0.874 RJupiter, and masses of 0.07
and 0.57 MJupiter (Huber et al. 2013). The RV semi-
amplitude is too small to be detectable with APOGEE,
so this target (H = 10.33) is an opportunity to test the
long-term RV precision level for our KOI program. Fig.
8 shows that the RV rms about the mean is 78 m s−1
over the entire baseline, in support of our stated goal to
achieve a long-term (1-2 year), relative RV precision of
∼ 100 m s−1. The APOGEE RVs are reported in Table
1.
3.1.5. SDSS-III KOI Program: KIC 6867766
KIC 6867766 (KOI 1798, Kp = 14.38, H = 12.99)
was listed in the Q1-Q6 KOI catalog as an exoplanet
candidate but has since been listed as "Not Disposi-
tioned" in later catalogs. This KOI is also listed in the
EB catalog. A shallow, 0.3% transit signal is present
with no obvious secondary feature at an orbital period
of 12.964725 days. The estimated radius from the KOI
catalog is 9.65 ± 1.5 R⊕. This target was observed a to-
tal of 25 times with APOGEE as part of our Kepler KOI
program within SDSS-III. We found the best 1D CCF
template match using a Teff = 5500 K, solar-metallicity
dwarf rotationally broadened to 14 km s−1. Upon visual
inspection, some of the CCFs were observed to be asym-
metric, and in some cases evidence of a double-peaked
0.0
−100
−50
0
RV (k /s)
50
100
Figure 9. Phase-folded 1D CCF for KIC 6867766 as observed by
APOGEE for a subset of the epochs. The CCF is clearly double-
peaked, and the two CCF components have maximum separation
near phases φ ∼ 0.25 and φ ∼ 0.75: a clear indication that the
KOI planet signal is the result of a binary companion.
CCF were present. In these situations, the parameters
of our best 1D CCF template are generally not reflective
of any component in the system: a multi-dimensional
cross-correlation analysis is required.
To test whether the Kepler transits might be due to
the binary showing up in the CCF, we calculate the 1D
CCFs for each APOGEE observation, normalize each
CCF to a peak value of unity, and then sort them based
on the orbital phase corresponding to each observation,
phase-folding on the time of transit and orbital period
as reported in the KOI catalog. We show only a few of
the CCFs sampling different orbital phases in Fig. 9 for
clarity. We find that this technique is quite effective at
finding time-variable CCF changes indicative of blended
SB2s.
In the case of KIC 6867766, the double-peaked
nature of the CCF is readily apparent, and phase-folds
nicely with the KOI orbital solution (maximum separa-
tion between CCF components near phases φ ∼ 0.25 and
φ ∼ 0.75 after phase-folding on KOI T0, blended compo-
nents near φ ∼ 0 and φ ∼ 0.5).
Upon further analysis, we determined that this sys-
tem includes three stellar mass objects: a late F-type
dwarf (KIC6867766A) with an early M-dwarf companion
(KIC6867766B) in a ∼ 13 day orbit, and a late G-type
dwarf (KIC6867766C) with no discernible velocity mo-
tion. We analyzed the APOGEE spectra with the TRI-
COR algorithm (Zucker et al. 1995), which uses a 3D
cross-correlation to derive the RV of three blended spec-
tral components and their relative flux ratios. Our anal-
ysis used spectral templates constructed from BT Settl
models (Allard et al. 2011), convolved to the APOGEE
spectral resolution and sampling. We optimized the tem-
plate stellar parameters to best match the observed spec-
tra, using the peak correlation power from our highest
S/N APOGEE spectra to assess the template match. All
three components used templates with solar metallicity
and log g = 4.5. The effective temperatures and rota-
tional velocities were: A: 6200 K, 10 km s−1; B: 3500 K,
APOGEE Spectroscopy of the Kepler Field
9
KIC6867766
A
C
B
q = 0.388± 0.006
100
50
0
−50
−100
4.0
2.0
0.0
−2.0
−4.0
20.0
10.0
0.0
−10.0
−20.0
)
1
−
s
m
k
(
y
t
i
c
o
e
V
l
i
l
a
d
a
R
A
C
−
A
O
)
1
−
s
m
k
(
B
C
−
B
O
)
1
−
s
m
k
(
0.0
0.2
0.4
0.6
Orbital Phase
0.8
1.0
Figure 10. Three-component RV solution for KIC 6867766 using
APOGEE RVs. Components A and B are an F+M EB pair that
produces the Kepler transit signal, while Component C is a (likely)
bound G dwarf companion with a long orbital period that dilutes
the A+B eclipse.
3 km s−1; C: 5200 K, 3 km s−1. We fixed the relative
H band flux ratios to be C/A = 0.35 and B/A = 0.05,
which are consistent with both the optimal flux ratios de-
rived by TRICOR and the physical flux ratios expected
for these stars.
Table 1 contains the RVs we derived for each of the
stellar components, while Table 2 lists the best (least-
squares derived) spectroscopic orbital parameters for the
13-day binary. We rejected some or all of the component
RVs measured in five of the epochs due to peak pulling
between the three components (generally A and C, but
occasionally all three), and these are indicated in Ta-
ble 1 with ellipses. Most of our APOGEE spectra have
median S/N of ∼ 25 − 30 per pixel, but four (2014-04-
11, 2014-04-14, 2014-05-11, 2014-06-18) had S/N . 10.
We recovered the M-dwarf signal in each of these low
S/N spectra and so retained them in our final orbital
solution, albeit with larger RV uncertainties that reflect
the poorer data quality. Fig. 10 shows the phase-folded
RV curves for the AB binary, and residuals from the de-
rived orbital solution. We also plot the measured RVs
for KIC6867766C, which are flat and very close to the
systemic velocity of the AB binary. This suggests that
KIC6867766C is a bound companion to the AB binary
in a very long-period orbit. We did not detect significant
change in the RV of KIC6867766C, so cannot estimate
its orbital period. We conclude that the KIC6867766
is a likely hierarchical triple system, and an exoplanet
false positive corresponding to Scenario 2 in Figure 1
(although in this case, the tertiary star is also within the
spectrograph's FoV).
3.2. Survey Efficiency
Several high-precision RV instruments are available in
the northern hemisphere to confirm Kepler exoplanet
candidates by achieving RV precisions of a few m s−1,
including HARPS-North, Keck HIRES, SOPHIE, and
HET HRS. These instruments are clearly capable of con-
ducting a reconnaissance survey of Kepler KPCs for false
positives, but their telescope time is most effectively
spent observing robust exoplanet candidates for con-
firmation and characterization purposes. Smaller tele-
scopes have been used to measure RVs for hundreds of
KOIs through the Kepler CFOP (e.g., McDonald 2.7m,
Tillinghast 1.5m), and while they have helped identify
the best candidates for higher precision RV follow-up,
they are limited to a single target at a time: most KOIs
have just one or two observations from these facilities. To
compare the efficiency of APOGEE against the larger-
aperture telescopes mentioned above, we calculate the
integration time per target required to achieve a photon-
limited RV precision of 100 m s−1. A quality factor Q is
calculated following Bouchy et al.
(2001), adapting in-
strument parameters summarized in Table 3. The qual-
ity factor represents the fundamental RV information
content of a spectrum, which depends on the number,
depths, and widths of spectral features. The Q values
are calculated using BT Settl stellar models for a range
of stellar effective temperatures (Teff), adopting a surface
gravity log g = 5.0, a solar metallicity, and no rotational
broadening.
Including a rotational broadening will af-
fect higher resolution instruments the most, resulting in
lower Q values.
From these Q values, it is then possible to calculate
the required integration time per object (an "effective
integration time") to achieve a given, photon-limited RV
precision (here taken to be 100 m s−1), via:
(cid:16)
c
σrv Q √πR2ǫF (cid:17)2
ntargets
t =
+ tover
(1)
where t is the effective integration time, c is the speed
of light, σrv is the desired RV precision, Q is the qual-
ity factor, R is the telescope's effective aperture radius,
ǫ is the total throughput (as a percentage) of the tele-
scope and instrument, F is the flux in photons per second
per unit area, tover is the overhead per integration, and
ntargets is the number of targets observed per integration.
Since we are operating in the photon-limited case, we do
not include readout noise for the instruments, nor con-
sider sources of systematic uncertainties such as residual
moonlight contamination (worse in the optical), or resid-
ual telluric lines and sky emission lines (worse in the NIR)
-- our interest is in calculating the photon limited case for
a direct comparison.
The fluxes are calculated from the BT Settl models
after convolving with the appropriate filter transmis-
sion function: Johnson/Bessell V 5 for HARPS-North,
Keck HIRES, SOPHIE, and HET HRS; 2MASS H 6 for
APOGEE. The model fluxes are scaled to the zero mag-
nitude level using zero-level fluxes from Bessell et al.
(1998) in V and Cohen et al.
(2003) in 2MASS, which
are then further scaled to a desired apparent magnitude.
We calculate the effective integration time for each in-
strument by including an estimated overhead for detector
readout, telescope slew, target acquisition and calibra-
tions. Since overhead times between integrations depend
5
http://www.ctio.noao.edu/~points/SOIFILTERS/filters/
maintext.html
6 http://www.ipac.caltech.edu/2mass/releases/allsky/doc/
sec3_1b1.html
10
Scott W. Fleming et al.
Figure 11. Total time (integration + overhead) required to
achieve a photon-limited RV precision of 100 m s−1 on a per target
basis, as a function of stellar Teff . Note that at this scale, HARPS-
North and Keck HIRES are nearly indistinguishable. APOGEE's
multiplexing capability is the driving factor in reducing the time
per target here.
on a variety of factors, we adopt three minutes as an
average overhead time for HARPS-North, Keck HIRES,
SOPHIE and HET HRS. For APOGEE, we use a mini-
mum integration time of 66.7 minutes and an overhead
time of 15 minutes, equivalent to the current survey's
integration time per visit for each field and its average
overhead time between fields. We do not reduce the in-
tegration time for APOGEE below this minimum value
because the total integration time is not based on any
one target's brightness, but rather an overall integration
time required for all targets in a field.
We calculate the required integration times t at a spe-
cific H (and corresponding V ) magnitude for both the
single object and multi-object instruments, to allow for
more direct comparisons. We use the median KPC host
star flux level of H ∼ 13.5, along with the appropriately
scaled V fluxes for the optical instruments. The average
number of KPC host stars within each APOGEE field
that have H < 13.5 is 89, so we use this as the number
of KPC hosts that can be observed simultaneously with
APOGEE's multiplexing capability for the purposes of
comparing against the single-object instruments. All of
these input parameters are summarized in Table 3.
Fig. 11 displays the required integration times per tar-
get to achieve a photon-limited RV precision of 100 m s−1
for a variety of spectral types. At this RV precision,
APOGEE is approximately three times as efficient on
a per target basis, primarily due to APOGEE's abil-
ity to observe multiple KPC hosts simultaneously, and
the fact that at this RV precision, the other instruments
are dominated by overheads per target. As the targeted
precision level increases, the other instruments become
increasingly more efficient compared to APOGEE, re-
flective of the fact that they are no longer dominated by
overheads. These basic calculations serve to demonstrate
how the multiplexing capability of APOGEE enables an
efficient survey at modest RV precision compared to the
single-object instruments, and that the telescope time
for those instruments is best spent on achieving high RV
precision to confirm new exoplanets.
Fig. 12 plots the orbital period versus H magnitude for
Figure 12. Orbital period versus H magnitude for the current
KPCs. The dashed, red lines demarcate H = 14 and P = 100 days.
Host stars brighter than this H limit and candidates with periods
below 100 days represent > 80% of the current KPC catalog, and
are particularly well-suited for APOGEE to characterize the orbits
of binaries. Our SDSS-III and SDSS-IV KOI programs observe
targets over more than one year baselines, however, and can detect
stellar companions out to the longest KOI orbital periods.
the current KPC catalog. Host stars with H < 14 and
P < 100 days are particularly well-suited for APOGEE
to characterize any binary star orbits, and represent more
than 80% of the current KPCs. Note, however, that even
KPCs with longer orbital periods, extending to at least
a few hundred days, can be identified as binaries even if
the observing baseline does not cover the entire orbital
period. If a more conservative limit of H < 13 is applied,
more than 47% of the KPCs would be included. Scenario
5 in Fig. 1 relies most heavily on achieving high signal-
to-noise ratio observations, and thus has the brightest
limiting magnitude within a survey; however, it is also
the false positive scenario for which APOGEE is least
sensitive.
4. DISCUSSION AND COMPARISON TO PRIOR WORK
4.1. Comparison With Other Efforts
A variety of techniques exist to determine whether a
given KPC might be a false positive using only Kepler
photometric information. Examination of the transits
can be done to ensure that the odd/even transits have
the same depth, that there are no ellipsoidal variations,
and that the positions of the flux centroids do not vary
with brightness changes, all of which are signs that the
transit event may be due to a diluted EB (Steffen et al.
2010; Torres et al. 2011). Another method is to gen-
erate a large grid of synthetic EB blend scenarios and
compare the models to the observed lightcurves via a
χ2 analysis (e.g., BLENDER, Torres et al. 2011), al-
though the technique can be computationally expensive
and difficult to apply to all KPCs en masse. In addition,
imaging surveys (e.g., Howell et al. 2011; Adams et al.
2012) can be used to inform the photometric analyses de-
scribed above, particularly to identify fainter stars that
exist within the lightcurve aperture, and to search for
wide stellar companions to study any relationships be-
tween exoplanet properties and host star multiplicity
(Wang et al. 2014).
Another technique makes use of stellar population
APOGEE Spectroscopy of the Kepler Field
11
synthesis and Galactic models to estimate the prob-
ability that a given transit signal
is due to an EB.
Morton & Johnson (2011) used such a technique to esti-
mate the false positive probability (FPP) of KPCs, and
concluded that 90% of the KPCs had FPPs < 10%, a
result that was used by other authors in subsequent sta-
tistical analyses of the KPC planet candidates. However,
the technique presented in Morton & Johnson (2011) re-
lied on KPCs to be vetted to the fullest extent possible
using the Kepler photometry, specifically, the removal
of V-shaped transits and searches for faint secondary
eclipses, whereas the table of KPCs presented in their
paper was not limited to these pre-vetted KPCs.
An updated version of the technique by Morton (2012)
accounted for more false positive scenarios and clarified
the importance of pre-vetting KPCs before performing
the statistical analysis. The updated implementation
can be run fairly quickly (of order 10 minutes per star),
and has been verified by testing it on confirmed KPCs
and known false positive KPCs; however, the technique
works best with additional observations (imaging to de-
tect close, visual companions and at least one high reso-
lution spectrum to get coarse stellar parameters). In ad-
dition, the Bayesian modeling is dependent on a variety
of model assumptions regarding Galactic structure, stel-
lar population synthesis, distribution of binary proper-
ties, and the frequency of exoplanets for various types of
stars. While the framework explicitly accounts for such
assumptions through the adopted priors, and is fairly
trivial to update when new knowledge is obtained about
any of these distributions, direct spectroscopic or photo-
metric observations of false positives are the least model-
dependent approach to derive the false positive statistics
of KPCs.
The Kepler CFOP program has been conducting
spectroscopic and imaging campaigns to identify blend
sources of KOIs, and, as previously mentioned, have col-
lected thousands of RV measurements for hundreds of
KOIs, in addition to high-resolution imaging to search for
faint companions unresolved in the Kepler photometric
aperture. Our program selected targets independently
of those observed by the CFOP program, since our goal
was to minimize selection bias in the KOIs we observe.
Our program complements the CFOP in a variety of
ways. The CFOP imaging data can identify wide stellar
companions, while our RV measurements over 18-month
baselines can detect linear RV trends of intermediate-
period companions that are unresolved by AO or lucky
imaging. The abundances of the ∼15 elements within
APOGEE's H band spectral window can be compared
and contrasted with elements accessible in the CFOP
optical spectra.
Many of the CFOP spectroscopic observations have ob-
tained just one or two spectra near predicted quadrature
phases, where any contaminating spectral lines from a
stellar companion are at maximum separation. This is of-
ten sufficient to identify a subset of false positive scenar-
ios: eclipsing stellar companions orbiting the KIC star,
for example.
In contrast, each of our KOI targets in
SDSS-III/IV (except those few that were part of our Ke-
pler EB program) obtain more than twenty APOGEE
RVs, which allow us to fully characterize the orbits of
any bound companions causing the false positive transit
signals.
In the case of eclipsing low-mass secondaries,
this further enables a study of the fundamental mass-
radius relationship for K and M dwarfs, since precise
mass ratios at the 1% level are achievable. Our mul-
tiple RV measurements can also be used to search for
(and place limits on) the presence of any longer-period,
non-transiting companions (stellar or otherwise), for the
study of multiplicity amongst Kepler planet hosts.
Studies have found that the false positive rates for
various subsets of KPCs are larger than the ones found
from the Morton & Johnson (2011) study, whose quoted
statistics are only valid for fully pre-vetted KPCs, and
that this rate might differ depending on the orbital pe-
riod and transit depth of the KPCs. Col´on et al. (2012)
made use of multi-color differential photometry to test
for false positives due to diluted EBs whose components
have sufficiently different colors. They observed a total
of four KPCs that had short periods (P < 6 days) and
small radii (Rp < 5 R⊕), and found evidence that two
of the four were likely due to diluted EBs, excluding an
overall false positive rate of 10% with 99% confidence.
Santerne et al. (2012) collected spectroscopic RVs of 33
giant planet KPCs using the SOPHIE spectrograph and
found a false positive probability of at least 35% within
their sample, where a majority of false positives were due
to EBs. Their sample size of 33 KPCs was partly limited
by their telescope resource: a single-object spectrograph
observing in the optical using a 2m-class telescope, cor-
responding to an effective magnitude limit in the Kepler
bandpass of Kp . 14.7. This magnitude limit removes
almost half of the total KPCs. Utilizing a multi-object,
NIR spectrograph, such as APOGEE, increases the rate
of data collection while also increasing the total number
of KPCs able to be observed.
4.2. Abundances of KOI Host Stars
Beyond identifying false positive KPCs as binaries,
a variety of additional science projects can be done
with the NIR APOGEE spectra. One such example
is the study of chemical abundance patterns in planet
host stars compared to stars not known to host plan-
ets. One of APOGEE's primary goals is to measure the
chemical abundances of many elements to study stel-
lar populations within the Milky Way. These abun-
dances are measured using the APOGEE Stellar Pa-
rameters and Chemical Abundances Pipeline (ASPCAP,
2015), which consists of a suite
Garc´ıa P´erez et al.
of software codes to analyze,
in an automated fash-
ion, the APOGEE spectra. The main component of
the code is FERRE7, a Fortran optimization code that
searches for the set of parameters that best match
each APOGEE spectrum. FERRE was originally devel-
oped in the context of low-resolution SDSS spectroscopy
(Allende Prieto 2004; Allende Prieto et al. 2006), and
has subsequently evolved and been used in other con-
texts (Allende Prieto et al. 2008a, 2009; Brown et al.
2012; Kilic et al. 2012). The APOGEE band (1.5-1.7
µm) is rich in transitions from many elements in cool
stars. Abundances for 15 elements can be derived from
sufficiently high resolution (R > 20000) and signal-to-
noise per pixel (S/N> 100) spectra in this spectral win-
dow: C, N, O, Na, Mg, Al, Si, K, Ca, Ti, V, Mn, Fe, Ni,
Cr. A S/N level approaching 100 per pixel is expected
7 FERRE is available at http://hebe.as.utexas.edu/ferre
12
Scott W. Fleming et al.
versus Neptunian-mass planets is 0.20 dex, which is sig-
nificant and indicates that the metallicity populations for
stars with Jovian-mass planets are not the same as those
which host the smaller Neptunian-mass planets. Oth-
ers have used the Kepler sample to extend the planet-
metallicity correlation down to terrestrial-sized planets
(Buchhave et al. 2012), and have found that terrestrial-
sized planets fall into well-defined host-star metallicity
regimes (Buchhave et al. 2014). These results suggest
that metallicity may also influence the distribution of
planetary masses within extrasolar systems.
In addi-
tion to stellar metallicity, there are also suggestions that
stellar mass plays a role, such that the dominant plane-
tary mass decreases as the parent star's mass decreases
(Ghezzi et al. 2010; Mayor et al. 2011), at least for
main sequence stars. While massive subgiant and giant
stars show trends similar to main sequence stars, low-
mass giants show a different behavior (Maldonado et al.
2013). It is likely that some combination of stellar mass
and metallicity influences the type of planetary system
that will form (Johnson et al. 2010). APOGEE can
provide a statistically significant control sample for such
studies.
In addition to overall metallicity and stellar mass play-
ing a role, the detailed chemistry of the parent cloud in
which the system forms may also hold clues to further
understanding planet formation. Recent findings sug-
gest that specific abundance patterns, such as Mg/Fe,
may influence the likelihood that a star hosts an under-
lying planetary system (Adibekyan et al. 2012a), and
that enhancement in alpha elements may favor the for-
mation of rocky planets, even for stars with low iron
abundances (Adibekyan et al. 2012b). The C/O ratio
in the parent cloud is also found in some studies to be en-
riched in planet hosting systems, with C/O ratios > 0.8
(Delgado Mena et al. 2010; Petigura & Marcy 2011),
however, these results have been questioned by Fortney
(2012), and have not been confirmed by other groups
(Nissen 2013; Teske et al. 2014). Brugamyer et al.
(2011) find that the silicon abundance (and not the
oxygen abundance) is a key element, as they find that
their planet detection rate depends strongly on the sil-
icon abundance of the host star. A difference in the
Si abundance is also found for the XO-2 binary host
stars, where XO-2N is found to be enhanced relative
to XO-2S (Teske et al. 2015). Mel´endez et al.
(2009),
Gonz´alez Hern´andez et al.
(2010), and Schuler et al.
(2011) present intriguing results suggesting that low-
amplitude chemical signatures point to selective accre-
tion or depletion of refractory and volatile elements in
stellar exoplanetary hosts. In particular, trends of abun-
dances with condensation temperature (Tcond) are used
as diagnostics, and these can be defined from the abun-
dances of the 15 chemical elements covered by APOGEE,
which include C, N, O (volatiles) and Si, Ti and Al (re-
fractories). The investigation of such trends in samples of
Kepler stars with confirmed planets of different masses,
and including the smallest planets to date, provides an
unprecedented database in order to probe the importance
of Tcond trends in this context.
5. SUMMARY
In this paper, we highlight the importance of an RV
survey of KPCs to better determine the false positive
Figure 13. Comparison of [M/H] values from ASPCAP DR12 and
Buchhave et al. (2014) for stars in common. The ASPCAP values
are taken from the '"uncalibrated" (initial fit) values ("FPARAM"
array). We find that a majority of these targets fall within 0.1 dex
of the Buchhave et al.
(2014) values (grey dashed lines). These
boundaries also represent the typical 1σ uncertainties for both sets
of values. There are not enough targets with "externally cali-
brated" values ("PARAM" array) in ASPCAP DR12 to make a
statement about how those compare to Buchhave et al. (2014).
to be achieved out to H ∼ 14 by coadding the multiple
visits.
A detailed analysis of elemental abundances for KOIs
with and without exoplanets is beyond the scope of this
introductory paper. To get some sense of the metal-
licites coming from the automated ASPCAP pipeline,
we have compared the "uncalibrated" [M/H] values from
ASPCAP in DR12 with those found in Buchhave et al.
(2014) for stars in common. The "uncalibrated" values
are the parameters that come from the initial fit. A
subset of targets have additional, external calibrations
applied to their stellar parameters (such as metallici-
ties of clusters from the literature). We refer the reader
to the ASPCAP DR12 documentation8 for full details.
There are not enough targets that have had external
calibrations applied to make any statement regarding
their agreement. However, we find a total of 128 KOI
host stars observed in common that have "uncalibrated"
[M/H] values in DR12. We find the agreement to be
promising (Fig. 13): a majority of targets agree within
0.1 dex (grey lines in Fig. 13), despite the fact that AS-
PCAP has been calibrated primarily to work on bright
giants. We have not made any cuts based on ASPCAP
processing flags; this is a comparison using DR12 [M/H]
values "as they are" versus the Buchhave et al.
(2014)
values. As such, the relation in Fig. 13 should be consid-
ered as preliminary, and likely to be improved upon in
future analyses.
There are tantalizing hints for different heavy ele-
ment patterns in planet hosts relative to the field that
could be induced by preferential removal of heavy ele-
ments in the disk.
It's been shown that stars hosting
Jovian-mass planets tend to be more metal-rich than
stars with only Neptunian-mass planets (Sousa et al.
2008; Ghezzi et al. 2010; Sousa et al. 2011). The over-
all shift in values of [Fe/H] between stars with Jovian-
8 http://www.sdss.org/dr12/irspec/parameters/
APOGEE Spectroscopy of the Kepler Field
13
rate, and demonstrate that APOGEE can efficiently con-
duct such a survey of KPCs to identify, and in many
cases characterize the orbits of, false positive KPCs. We
have shown that the APOGEE instrument is capable of
achieving an RV precision of ∼ 100 m s−1 using observa-
tions of the confirmed exoplanet host star KIC 6448890,
as well as the Kepler EB KIC 01571511, which produces
planet-sized eclipses and has an H magnitude similar to
many KPCs. We find that the transit signal of KIC
3848972 is not caused by a blue, stellar secondary orbit-
ing the primary star, and do not find any evidence of a
faint, red companion in Keck AO images that could be
the source of the Kepler transit events. Further inves-
tigation is merited before the true nature of this KOI
can be confidently identified. We find HET RV varia-
tions that phase to the KOI period and ephemeris for
KIC 3861595, but our APOGEE RVs are inconsistent
with the HET RV semiamplitude, and we find evidence
of line bisector variations. This target was part of our
EB program in SDSS, so we only have a few spectra
to work with, and the RV uncertainties have not been
fully vetted for rapdily rotating stars such as this one.
As such, we can not definitively determine the nature of
this KPC, and urge more spectra be obtained to exam-
ine both the RV and line bisector variations. Finally, we
find that KIC 6867766 is a triple system, composed of
an F+M EB and a wide, bound G dwarf tertiary. The
F+M EB phases to the KOI period and ephemeris, and
the diluted eclipses are the source of the KPC transits.
As such, we can confidently identify this KOI as a false
positive exoplanet candidate.
Not only can the data from such a survey be used to de-
termine the false positive rate of KPCs and vet the sam-
ple to identify the best candidates for high-precision RV
observations, but it will enable ancillary science projects
in fundamental stellar astrophysics though observations
of EBs, studies of intrinsically rare short-period com-
panions (such as brown dwarfs), and detailed chemical
abundances of exoplanet host stars. At the precision
level of 100 m s−1, APOGEE is a more efficient instru-
ment compared to HARPS-North, Keck HIRES, and SO-
PHIE, due to its multiplexing capability and because the
single-object spectrographs are dominated by overheads.
Our survey to detect false positives refines the target
selection for higher precision RV instruments, enabling
them to focus on the best exoplanet candidates. It will
allow for improved statistical studies of the Kepler exo-
planet population by determining the false positive rate
of KPCs due to physically-bound binaries, as well as any
trends in the false positive rate with orbital period or
stellar properties.
We thank the anonymous referee for detailed com-
ments that improved the quality of this publication, es-
pecially their suggestion to conduct more analysis on
the KOI-4 RVs. We acknowledge support from NSF
award AST1006676, AST 1126413, and AST 1310885.
This work was partially supported by funding from the
Center for Exoplanets and Habitable Worlds. The Cen-
ter for Exoplanets and Habitable Worlds is supported
by the Pennsylvania State University, the Eberly Col-
lege of Science, and the Pennsylvania Space Grant Con-
sortium. CAP acknowledges funding from the Spanish
Ministry of Economy and Competitiveness (MINECO)
through grant AYA2011-26244. This research has made
use of the SIMBAD database, operated at CDS, Stras-
bourg, France. This publication makes use of data prod-
ucts from the Two Micron All Sky Survey, which is a
joint project of the University of Massachusetts and the
Infrared Processing and Analysis Center/California In-
stitute of Technology, funded by the National Aeronau-
tics and Space Administration and the National Science
Foundation. Some of the data presented in this paper
were obtained from the Mikulski Archive for Space Tele-
scopes (MAST). STScI is operated by the Association
of Universities for Research in Astronomy, Inc., under
NASA contract NAS5-26555. Support for MAST for
non-HST data is provided by the NASA Office of Space
Science via grant NNX13AC07G and by other grants and
contracts. This paper includes data collected by the Ke-
pler mission. Funding for the Kepler mission is provided
by the NASA Science Mission directorate. This research
has made use of the NASA Exoplanet Archive, which is
operated by the California Institute of Technology, under
contract with the National Aeronautics and Space Ad-
ministration under the Exoplanet Exploration Program.
This work was based on observations with the SDSS
2.5-meter telescope. Funding for SDSS-III has been pro-
vided by the Alfred P. Sloan Foundation, the Participat-
ing Institutions, the National Science Foundation, and
the U.S. Department of Energy Office of Science. The
SDSS-III web site is http://www.sdss3.org/. SDSS-
III is managed by the Astrophysical Research Consor-
tium for the Participating Institutions of the SDSS-
III Collaboration including the University of Arizona,
the Brazilian Participation Group, Brookhaven National
Laboratory, University of Cambridge, Carnegie Mellon
University, University of Florida, the French Participa-
tion Group, the German Participation Group, Harvard
University, the Instituto de Astrofisica de Canarias, the
Michigan State/Notre Dame/JINA Participation Group,
Johns Hopkins University, Lawrence Berkeley National
Laboratory, Max Planck Institute for Astrophysics, Max
Planck Institute for Extraterrestrial Physics, New Mex-
ico State University, New York University, Ohio State
University, Pennsylvania State University, University of
Portsmouth, Princeton University, the Spanish Partici-
pation Group, University of Tokyo, University of Utah,
Vanderbilt University, University of Virginia, University
of Washington, and Yale University.
The parameters used to calculate the background EB blend probability of KIC 3861595 are summarized in Table 4.
APPENDIX
TRILEGAL INPUT PARAMETERS
Adams, E. R., Ciardi, D. R., Dupree, A. K., et al. 2012, AJ, 144, 42
REFERENCES
14
Scott W. Fleming et al.
Adibekyan, V. Z., Santos, N. C., Sousa, S. G., et al. 2012, A&A, 543, A89
Adibekyan, V. Z., Delgado Mena, E., Sousa, S. G., et al. 2012, arXiv:1209.6272
Allard, F., Homeier, D., & Freytag, B. 2011, 16th Cambridge Workshop on Cool Stars, Stellar Systems, and the Sun, 448, 91
Allende Prieto, C. 2004, Astronomische Nachrichten, 325, 604
Allende Prieto, C., Beers, T. C., Wilhelm, R., et al. 2006, ApJ, 636, 804
Allende Prieto, C., Sivarani, T., Beers, T. C., et al. 2008, AJ, 136, 2070
Allende Prieto, C., Majewski, S. R., Schiavon, R., et al. 2008, Astronomische Nachrichten, 329, 1018
Allende Prieto, C., Hubeny, I., & Smith, J. A. 2009, MNRAS, 396, 759
Batalha, N. M., Rowe, J. F., Bryson, S. T., et al. 2013, ApJS, 204, 24
Bender, C. F., & Simon, M. 2008, ApJ, 689, 416
Bender, C. F., Mahadevan, S., Deshpande, R., et al. 2012, ApJ, 751, L31
Bessell, M. S., Castelli, F., & Plez, B. 1998, A&A, 333, 231
Borucki, W. J., Koch, D., Basri, G., et al. 2010, Science, 327, 977
Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 728, 117
Borucki, W. J., Koch, D. G., Basri, G., et al. 2011, ApJ, 736, 19
Bouchy, F., Pepe, F., & Queloz, D. 2001, A&A, 374, 733
Brown, T. M., Latham, D. W., Everett, M. E., & Esquerdo, G. A. 2011, AJ, 142, 112
Brown, W. R., Kilic, M., Allende Prieto, C., & Kenyon, S. J. 2012, ApJ, 744, 142
Brugamyer, E., Dodson-Robinson, S. E., Cochran, W. D., & Sneden, C. 2011, ApJ, 738, 97
Buchhave, L. A., Latham, D. W., Johansen, A., et al. 2012, Nature, 486, 375
Buchhave, L. A., Bizzarro, M., Latham, D. W., et al. 2014, Nature, 509, 593
Butler, R. P., Marcy, G. W., Williams, E., et al. 1996, PASP, 108, 500
Caldwell, D. A., Kolodziejczak, J. J., Van Cleve, J. E., et al. 2010, ApJ, 713, L92
Carter, J. A., Fabrycky, D. C., Ragozzine, D., et al. 2011, Science, 331, 562
Chaplin, W. J., Kjeldsen, H., Christensen-Dalsgaard, J., et al. 2011, Science, 332, 213
Christiansen, J. L., Jenkins, J. M., Barclay, T. S., et al. 2012, arXiv:1208.0595
Cohen, M., Wheaton, W. A., & Megeath, S. T. 2003, AJ, 126, 1090
Col´on, K. D., Ford, E. B., & Morehead, R. C. 2012, arXiv:1207.2481
Coughlin, J. L., L´opez-Morales, M., Harrison, T. E., Ule, N., & Hoffman, D. I. 2011, AJ, 141, 78
Delgado Mena, E., Israelian, G., Gonz´alez Hern´andez, J. I., et al. 2010, ApJ, 725, 2349
Dotter, A., Chaboyer, B., Jevremovi´c, D., et al. 2008, ApJS, 178, 89
Eisenstein, D. J., Weinberg, D. H., Agol, E., et al. 2011, AJ, 142, 72
Ford, E. B., Ragozzine, D., Rowe, J. F., et al. 2012, ApJ, 756, 185
Fortney, J. J. 2012, ApJ, 747, L27
Fukugita, M., Ichikawa, T., Gunn, J. E., et al. 1996, AJ, 111, 1748
Garc´ıa P´erez, A. E. et al. 2015, in prep
Ghezzi, L., Cunha, K., Smith, V. V., et al. 2010, ApJ, 720, 1290
Gilliland, R. L., Chaplin, W. J., Dunham, E. W., et al. 2011, ApJS, 197, 6
Girardi, L., Barbieri, M., Groenewegen, M. A. T., et al. 2012, Red Giants as Probes of the Structure and Evolution of the Milky Way, 165
Gonz´alez Hern´andez, J. I., Israelian, G., Santos, N. C., et al. 2010, ApJ, 720, 1592
Gunn, J. E., et al. 2006, AJ, 131, 2332
Hirano, T., Sanchis-Ojeda, R., Takeda, Y., et al. 2012, ApJ, 756, 66
Holman, M. J., Fabrycky, D. C., Ragozzine, D., et al. 2010, Science, 330, 51
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS, 201, 15
Howell, S. B., Everett, M. E., Sherry, W., Horch, E., & Ciardi, D. R. 2011, AJ, 142, 19
Huber, D., Carter, J. A., Barbieri, M., et al. 2013, Science, 342, 331
Huber, D., Silva Aguirre, V., Matthews, J. M., et al. 2014, ApJS, 211, 2
Kraus, A. L., & Hillenbrand, L. A. 2007, AJ, 134, 2340
Johnson, J. A., Aller, K. M., Howard, A. W., & Crepp, J. R. 2010, PASP, 122, 905
Johnson, J. A., Clanton, C., Howard, A. W., et al. 2011, ApJS, 197, 26
Kilic, M., Brown, W. R., Allende Prieto, C., et al. 2012, ApJ, 751, 141
Koch, D. G., Borucki, W. J., Basri, G., et al. 2010, ApJ, 713, L79
Lagrange, A.-M., Desort, M., Galland, F., Udry, S., & Mayor, M. 2009, A&A, 495, 335
Mahadevan, S., Deshpande, R., Bender, C., et al. 2015, in prep
Majewski, S. R., Wilson, J. C., Hearty, F., Schiavon, R. R., & Skrutskie, M. F. 2010, IAU Symposium, 265, 480
Majewski, S. R., et al. 2015, in prep
Maldonado, J., Villaver, E., & Eiroa, C. 2013, A&A, 554, AA84
Marcy, G. W. & Butler, R. P. 2000, PASP, 112, 137
Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Messenger, 114, 20
Mayor, M., Marmier, M., Lovis, C., et al. 2011, arXiv:1109.2497
Mel´endez, J., Asplund, M., Gustafsson, B., & Yong, D. 2009, ApJ, 704, L66
M´esz´aros, S., Allende Prieto, C., Edvardsson, B., et al. 2012, arXiv:1208.1916
Moorhead, A. V., Ford, E. B., Morehead, R. C., et al. 2011, ApJS, 197, 1
Morton, T. D. 2012, arXiv:1206.1568
Morton, T. D., & Johnson, J. A. 2011, ApJ, 738, 170
Nidever, D. L., Holtzman, J. A., Allende Prieto, C., et al. 2015, arXiv:1501.03742
Nissen, P. E. 2013, A&A, 552, AA73
Ofir, A., Gandolfi, D., Buchhave, L., et al. 2012, MNRAS, 423, L1
Perruchot, S., Kohler, D., Bouchy, F., et al. 2008, Proc. SPIE, 7014, 17
Petigura, E. A., & Marcy, G. W. 2011, ApJ, 735, 41
Prsa, A., Batalha, N., Slawson, R. W., et al. 2011, AJ, 141, 83
Rhodes, M.D. & Budding, E. 2014, Ap&SS
Sahlmann, J., et al. 2011, A&A, 525, 95
Santerne, A., D´ıaz, R. F., Moutou, C., et al. 2012, A&A, 545, AA76
Santerne, A., Fressin, F., D´ıaz, R. F., et al. 2013, A&A, 557, AA139
Schuler, S. C., Flateau, D., Cunha, K., et al. 2011, ApJ, 732, 55
Slawson, R. W., Prsa, A., Welsh, W. F., et al. 2011, AJ, 142, 160
APOGEE Spectroscopy of the Kepler Field
15
Sousa, S. G., Santos, N. C., Mayor, M., et al. 2008, A&A, 487, 373
Sousa, S. G., Santos, N. C., Israelian, G., Mayor, M., & Udry, S. 2011, A&A, 533, AA141
Steffen, J. H., Batalha, N. M., Borucki, W. J., et al. 2010, ApJ, 725, 1226
Steffen, J. H., Fabrycky, D. C., Agol, E., et al. 2013, MNRAS, 428, 1077
Szentgyorgyi, A. H., & Fur´esz, G. 2007, Revista Mexicana de Astronomia y Astrofisica Conference Series, 28, 129
Teske, J. K., Cunha, K., Smith, V. V., Schuler, S. C., & Griffith, C. A. 2014, ApJ, 788, 39
Teske, J. K., Ghezzi, L., Cunha, K., et al. 2015, arXiv:1501.02167
Torres, G., Fressin, F., Batalha, N. M., et al. 2011, ApJ, 727, 24
Tull, R. G. 1998, Proc. SPIE, 3355, 387
Twicken, J. D., Clarke, B. D., Bryson, S. T., et al. 2010, Proc. SPIE, 7740, 69
Vogt, S. S., Allen, S. L., Bigelow, B. C., et al. 1994, Proc. SPIE, 2198, 362
Wang, J., Fischer, D. A., Xie, J.-W., & Ciardi, D. R. 2014, ApJ, 791, 111
Wilson, J. C., Hearty, F., Skrutskie, M. F., et al. 2010, Proc. SPIE, 7735, 77351C
Wilson, J. C., Hearty, F., Skrutskie, M. F., et al. 2012, Proc. SPIE, 8446
Wizinowich, P. L., Acton, D. S., Lai, O., et al. 2000, Proc. SPIE, 4007, 2
Wright, J. T., & Howard, A. W. 2009, ApJS, 182, 205
Wright, J. T., Roy, A., Mahadevan, S., et al. 2013, ApJ, 770, 119
Youdin, A. N. 2011, ApJ, 742, 38
Zamora, O., Garc´ıa-Hern´andez, D. A., Allende Prieto, C., et al. 2015, in prep
Zasowski, G., Johnson, J. A., Frinchaboy, P. M., et al. 2013, AJ, 146, 81
Zucker, S., Torres, G., & Mazeh, T. 1995, ApJ, 452, 863
Zucker, S. 2003, MNRAS, 342, 1291
16
Scott W. Fleming et al.
RVs for KIC Stars - All RVs in km s−1
Table 1
KIC ID
BJD TDB
RVA
1σ
RVB
1σ
RVC
1σ
Instrument
1571511
1571511
1571511
3848972
3848972
3848972
3861595
3861595
3861595
3861595
3861595
3861595
3861595
3861595
3861595
3861595
3861595
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6448890
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
6867766
2455811.61304
2455840.59327
2455851.57845
2455811.61297
2455840.59327
2455851.57848
2455789.84195
2455796.82710
2455797.80608
2455801.80843
2455803.80347
2455813.70317
2455823.72718
2455840.66180
2455849.57900
2455851.64939
2455866.56998
2456368.99828
2456411.92027
2456557.73343
2456559.72336
2456560.72108
2456584.63225
2456585.63076
2456757.89294
2456758.90229
2456760.90571
2456761.87281
2456762.86860
2456763.88112
2456783.83567
2456784.82195
2456785.82543
2456786.79845
2456787.80934
2456788.84307
2456812.74509
2456814.75547
2456815.78552
2456816.76627
2456817.76198
2456818.76458
2456819.76222
2456820.75601
2456557.73337
2456559.72331
2456560.72103
2456584.63222
2456585.63072
2456757.89298
2456758.90233
2456760.90575
2456761.87284
2456762.86864
2456763.88116
2456783.83568
2456784.82197
2456785.82544
2456786.79846
2456787.80935
2456788.84309
2456812.74507
2456814.75546
2456815.78551
2456816.76626
2456817.76197
2456818.76456
2456819.76220
2456820.75599
-24.401
-26.348
-18.927
-19.943
-20.161
-19.641
-23.052
-23.827
-23.204
-23.194
-24.285
-22.543
-23.628
-22.930
-24.600
-21.337
-21.608
-55.391
-55.513
-55.598
-55.644
-55.571
-55.582
-55.644
-55.622
-55.801
-55.586
-55.573
-55.621
-55.581
-55.712
-55.781
-55.702
-55.590
-55.640
-55.679
-55.620
-55.615
-55.607
-55.710
-55.632
-55.609
-55.567
-55.666
· · ·
38.598
40.497
28.939
37.993
· · ·
1.277
-20.442
-17.315
-10.374
· · ·
15.664
· · ·
-12.960
-21.181
-20.018
-12.259
-20.349
-10.567
· · ·
16.157
27.340
37.006
40.742
38.303
0.153
0.115
0.105
0.157
0.153
0.150
0.091
0.075
0.081
0.098
0.103
0.566
0.469
0.541
0.434
0.433
0.461
0.117
0.105
0.103
0.106
0.104
0.104
0.105
0.107
0.142
0.112
0.139
0.111
0.109
0.112
0.133
0.108
0.113
0.107
0.118
0.111
0.114
0.107
0.119
0.109
0.110
0.109
0.108
· · ·
0.383
0.365
0.399
0.446
· · ·
1.679
0.482
1.389
0.432
· · ·
0.590
· · ·
0.422
0.542
0.398
1.175
0.501
0.455
· · ·
0.886
0.445
0.398
0.422
0.356
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
-60.127
-65.538
-36.948
-64.834
· · ·
46.983
88.944
79.738
66.784
28.563
-3.902
37.085
73.438
89.949
92.288
55.283
90.837
70.185
32.014
-7.116
-35.030
-59.225
-65.495
-57.968
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
2.882
6.014
3.811
4.138
· · ·
9.919
3.869
11.967
3.542
3.463
3.310
3.697
3.067
5.130
4.554
9.284
6.466
6.978
3.443
4.691
4.168
3.500
3.783
3.165
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
· · ·
6.928
8.042
7.788
8.956
· · ·
8.567
10.516
10.500
9.447
· · ·
11.547
· · ·
9.843
9.443
9.495
7.295
8.471
9.955
· · ·
9.461
11.971
8.936
9.396
8.577
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
HET
· · ·
HET
· · ·
HET
· · ·
HET
· · ·
HET
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
· · ·
APOGEE
0.664 APOGEE
0.653 APOGEE
0.780 APOGEE
0.754 APOGEE
· · ·
APOGEE
3.754 APOGEE
0.778 APOGEE
3.354 APOGEE
0.760 APOGEE
· · ·
APOGEE
1.018 APOGEE
· · ·
APOGEE
0.759 APOGEE
0.887 APOGEE
0.656 APOGEE
1.769 APOGEE
0.870 APOGEE
0.891 APOGEE
· · ·
APOGEE
1.656 APOGEE
1.059 APOGEE
0.702 APOGEE
0.780 APOGEE
0.668 APOGEE
APOGEE Spectroscopy of the Kepler Field
17
KIC 6867766 A+B Orbital Parameters
Table 2
Parameter
Value
1σ
P (days)
Tp
e
ω (deg)
KA (km s−1)
KB (km s−1)
γ (km s−1)
MB /MA
12.964712
2456746.58
0.0553
128.6
30.77
79.4
10.87
0.3877
(fixed at Kepler value)
0.21
0.0054
5.7
0.14
1.2
0.14
0.0060
Instrument Parameters For Q Factor Calculation
Table 3
Instrument
Resolution
HARPS-Northa
Keck HIRESb
SOPHIEc
HETd
APOGEEe
115000
55000
75000
30000
22500
λmin
(cid:0)A(cid:1)
3830
5000
3820
4076
15100
λmax
(cid:0)A(cid:1)
6930
6200
6930
7838
17000
Eff. Aperture Total Throughput Overhead # Targets / Obs.
Radius (m)
(%)
(min)
(H ∼ 13.5)
3.58
10.0
1.93
9.2
2.12
8
13
4
3
16
3
3
3
3
15
1
1
1
1
89
a Instrument parameters taken from http://www.tng.iac.es/instruments/harps/.
b Instrument specs as reported in Johnson et al.
inst/hires/throughput.pdf, minus an additional 30% loss due to absorption from the iodine cell.
c Instrument parameters taken from http://www.obs-hp.fr/guide/sophie/sophie-info.html.
d Instrument parameters are for the 316g5936 cross-disperser in R = 30000 mode using ThAr as a wavelength calibration.
Although most planet work has been done using the R = 60000 mode, this resolution is not required to achieve 100 m s−1
RV precision. Efficiency taken from the HRS exposure time calculator http://het.as.utexas.edu/HET/hetweb/Instruments/
HRS/exp/exp_calc.html, and is calculated at the center of the telescope's observability track.
e Instrument parameters taken from Wilson et al.
of the 2.5m diameter (Gunn et al. 2006).
(2011). Efficiency is taken to be 18% from http://www2.keck.hawaii.edu/
(2012). Effective aperture radius includes a 30% loss due to obstruction
18
Scott W. Fleming et al.
TRILEGAL Parameters Used In KIC 3861595 Background EB Blend Probability Calculation
Table 4
Parameter
Value
Distance modulus resolution of Galaxy components
IMF for single stars
Binary fraction
Binary mass ratios
Extinction model
Extinction model hz,dust
Extinction model hR,dust
Extinction calibration at infinity
1σ extinction dispersion
Solar Galactocentric radius R⊙
Solar height above the disk z⊙
Thin disk model
Thin disk z0
Thin disk t0
Thin disk α
Thin disk hR,d
Thin disk radial cutoffs
Thin disk Σd (⊙)
Thin disk SFR+AMR
Thick disk model
Thick disk hz,td
Thick disk hR,td
Thick disk radial cutoffs
Thick disk Ωtd (⊙)
Thick disk SFR+AMR
Halo model
Halo rh
Halo qh
Halo Ωh (⊙)
Halo SFR+AMR
Bulge model
Bulge am
Bulge a0
Bulge y/x axial ratio η
Bulge z/x axial ratio ξ
Bulge Sun-GC-bar angle φ0
Bulge Ωb (GC)
Bulge SFR+AMR
0.1 mag
Chabrier lognormal
0.3
0.7 to 0.1
Exponential disk of form exp (cid:0)− z /hz,dust(cid:1) ∗ exp (cid:0)−R/hR,dust(cid:1)
110 pc
100000 pc
0.0378
0.
8700 pc
24.2 pc
squared hyperbolic secant
94.6902 pc
5.55079E9 yr
1.6666
2913.36 pc
0, 15000 pc
55.4082 M⊙ pc−2
2-step SFR + Fuhrman's AMR + α enhancement with age(yr) = 0.735097t + 0
squared hyperbolic secant
800 pc
2394.07 pc
0, 15000 pc
0.001 M⊙ pc−3
11-12 Gyr const. SFR + Z=0.008 with σ[M/H] = 0.1 dex with age(yr) = t + 0
Oblate r1/4 spheroid
2698.93 pc
0.583063
0.000100397 M⊙ pc−3
12-13 Gyr + Ryan & Norris [M/H] distribution with age (yr) = t + 0
triaxial bulge
2500 pc
95 pc
0.68
0.31
15◦
406 M⊙ pc−3
10 Gyr, Zoccali et al. 2003 [M/H] + 0.3 dex with age(yr) = t − 2E9
|
1712.01042 | 2 | 1712 | 2018-02-13T20:54:51 | A Neptune-mass Free-floating Planet Candidate Discovered by Microlensing Surveys | [
"astro-ph.EP"
] | Current microlensing surveys are sensitive to free-floating planets down to Earth-mass objects. All published microlensing events attributed to unbound planets were identified based on their short timescale (below two days), but lacked an angular Einstein radius measurement (and hence lacked a significant constraint on the lens mass). Here, we present the discovery of a Neptune-mass free-floating planet candidate in the ultrashort ($t_{\rm E}=0.320\pm0.003$ days) microlensing event OGLE-2016-BLG-1540. The event exhibited strong finite-source effects, which allowed us to measure its angular Einstein radius of $\theta_{\rm E}=9.2\pm0.5\,\mu$as. There remains, however, a degeneracy between the lens mass and distance. The combination of the source proper motion and source-lens relative proper motion measurements favors a Neptune-mass lens located in the Galactic disk. However, we cannot rule out that the lens is a Saturn-mass object belonging to the bulge population. We exclude stellar companions up to 15 au. | astro-ph.EP | astro-ph |
Submitted to ApJ.
Preprint typeset using LATEX style emulateapj v. 01/23/15
A NEPTUNE-MASS FREE-FLOATING PLANET CANDIDATE DISCOVERED BY MICROLENSING SURVEYS
Przemek Mr´oz1†, Y.-H. Ryu2, J. Skowron1, A. Udalski1, A. Gould2,3,4,
M. K. Szyma´nski1, I. Soszy´nski1, R. Poleski3, P. Pietrukowicz1,
and
S. Koz(cid:32)lowski1, M. Pawlak1, K. Ulaczyk5
(The OGLE Collaboration),
and
M. D. Albrow6, S.-J. Chung2,7, Y. K. Jung8, C. Han9, K.-H. Hwang2, I.-G. Shin8, J. C. Yee8, W. Zhu10,
S.-M. Cha2,11, D.-J. Kim2, H.-W. Kim2, S.-L. Kim2,7, C.-U. Lee2,7, D.-J. Lee2, Y. Lee2,11, B.-G. Park2,7, and
R. W. Pogge3
(The KMTNet Collaboration)
1Warsaw University Observatory, Al. Ujazdowskie 4, 00-478 Warszawa, Poland
2Korea Astronomy and Space Science Institute, Daejon 34055, Republic of Korea
3Department of Astronomy, Ohio State University, 140 West 18th Avenue, Columbus, OH 43210, USA
4Max-Planck-Institute for Astronomy, Konigstuhl 17, D-69117 Heidelberg, Germany
5Department of Physics, University of Warwick, Coventry CV4 7AL, UK
6University of Canterbury, Department of Physics and Astronomy, Private Bag 4800, Christchurch 8020, New Zealand
7Korea University of Science and Technology, 217 Gajeong-ro, Yuseong-gu, Daejeon, 34113, Republic of Korea
8Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
9Department of Physics, Chungbuk National University, Cheongju 28644, Republic of Korea
10Canadian Institute for Theoretical Astrophysics, University of Toronto, Toronto, ON M5S 3H8, Canada and
11School of Space Research, Kyung Hee University, Yongin, Kyeonggi 17104, Republic of Korea
Submitted to ApJ.
ABSTRACT
Current microlensing surveys are sensitive to free-floating planets down to Earth-mass objects. All
published microlensing events attributed to unbound planets were identified based on their short
timescale (below two days), but lacked an angular Einstein radius measurement (and hence lacked a
significant constraint on the lens mass). Here, we present the discovery of a Neptune-mass free-floating
planet candidate in the ultrashort (tE = 0.320 ± 0.003 days) microlensing event OGLE-2016-BLG-
1540. The event exhibited strong finite-source effects, which allowed us to measure its angular Einstein
radius of θE = 9.2±0.5 µas. There remains, however, a degeneracy between the lens mass and distance.
The combination of the source proper motion and source-lens relative proper motion measurements
favors a Neptune-mass lens located in the Galactic disk. However, we cannot rule out that the lens is a
Saturn-mass object belonging to the bulge population. We exclude stellar companions up to ∼ 15 au.
Subject headings: planets and satellites: detection, gravitational lensing: micro
1.
INTRODUCTION
Gravitational microlensing enables detecting dark ob-
jects in a broad mass range: from black holes and neu-
tron stars to Earth-sized planets (Paczy´nski 1986).
In
particular, current microlensing surveys are sensitive to
free-floating planets, which are gravitationally unbound
to any star. A characteristic timescale of a microlensing
event (known as the Einstein time tE) depends on the
relative lens-source proper motion µrel and the angular
Einstein radius θE:
tE =
θE
µrel
=
√κM πrel
µrel
(1)
where M is the lens mass, πrel = 1 au (D−1
S ) is
the lens-source relative parallax, and κ = 4G/(c2 au) =
8.14 mas/M(cid:12). Here, DL and DS are distances to the lens
and source, respectively.
L − D−1
The Einstein timescale is the only physical parameter
that can be measured for the majority of microlensing
events. As the timescale is proportional to the square
† Corresponding author: [email protected]
root of mass, it is expected that events caused by free-
floating planets are very short (tE (cid:46) 2 days). However,
the mass measurement requires the knowledge of two ad-
ditional physical parameters: the angular Einstein radius
θE and the microlens parallax πE = πrel/θE.
Although the angular Einstein radius is routinely mea-
sured in binary microlensing events via the finite-source
effect (Udalski et al. 1994; Mao et al. 1994; Nemiroff
& Wickramasinghe 1994), such a measurement is much
harder for single lensing events because it requires that
the source passes almost exactly over the lens to pro-
duce a detectable finite-source signal (Alcock et al. 1997;
Yoo et al. 2004). It is expected that finite-source effects
should be strong for Earth-mass lenses, because the an-
gular size of the source is comparable to the Einstein ring
radius (Bennett & Rhie 1996; Ma et al. 2016). To date,
however, no such measurements have been reported.
The microlens parallax measurements are even harder
for free-floating planets. The subtle deviations from the
standard microlensing light curve due to parallax can be
detected in long-timescale events, as the Earth-based ob-
server moves along the orbit (Gould 1992). Parallax can
be also measured using simultaneous ground- and space-
based observations (Refsdal 1966), for example, with the
2
Mr´oz et al.
Spitzer satellite (Dong et al. 2007; Udalski et al. 2015b).
However, Spitzer operations require the targets to be up-
loaded to the spacecraft at least three days in advance,
making observations of short events nearly impossible.
The problem can be overcome with continuous, survey-
mode observations (e.g., Henderson & Shvartzvald 2016;
Gould 2016). Such an experiment was conducted dur-
ing the K2 Campaign 9 (Henderson et al. 2016; Penny
et al. 2017; Zhu et al. 2017), but owing to the difficul-
ties in extracting the photometry from crowded regions
of the Galactic bulge, no observations of short-timescale
microlensing events from K2 C9 were reported so far.
Information about the mass function of lenses, includ-
ing free-floating planets, can be inferred from a statistical
analysis of the distribution of timescales of a large sample
of microlensing events. The first such measurement was
attempted by Sumi et al. (2011), who analyzed a sample
of 474 microlensing events detected by the Microlensing
Observations in Astrophysics (MOA) group. They found
an excess of nine short2 events, relative to what was ex-
pected from brown dwarfs and stars, and they attributed
this excess to a large population of Jupiter-mass plan-
ets, which should be nearly twice as common as main-
sequence stars.
Clanton & Gaudi (2017) modeled the microlensing sig-
nal expected from exoplanets on wide orbits using con-
straints from microlensing, radial velocity, and direct
imaging surveys and concluded that at most ∼ 40% of
short-timescale events detected by Sumi et al. (2011) can
be interpreted as due to wide-orbit planets. However, the
statistical significance of Sumi et al.'s results is largely
based on three shortest-timescale events (tE < 1 days).
As we mentioned above, one measurement is incorrect
and the model by Clanton & Gaudi (2017) shows that
the remaining two are statistically consistent with being
wide-orbit planets. That model still cannot account for a
small overabundance of events with timescales between
1 -- 2 days (see Figures 4 and 5 from Clanton & Gaudi
2017), but the statistical significance of the remaining ex-
cess relative to the short-timescale events expected from
stars and brown dwarfs is small.
A large population of Jupiter-mass free-floating plan-
ets suggested by Sumi et al. (2011) was difficult to recon-
cile with censuses of substellar objects in young clusters
and star-forming regions and with predictions of planet-
formation theories. For example, Pena Ram´ırez et al.
(2012) and Scholz et al. (2012) analyzed substellar mass
functions of the young clusters σ Orionis and NGC 1333,
finding that free-floating planetary-mass objects are at
least an order of magnitude less common than main-
sequence stars. These observations are incomplete for
masses below ∼ 6 MJup, so direct comparisons with mi-
crolensing surveys are difficult. Several mechanisms of
free-floating planet production have been proposed (e.g.,
Veras & Raymond 2012), but none of them is capable of
explaining the large number of Jupiter-mass free-floaters
suggested by Sumi et al. On the other hand, Earth- and
super-Earth-mass planets can be scattered and ejected
much more efficiently (Pfyffer et al. 2015; Ma et al. 2016;
Barclay et al. 2017).
2 They reported ten events with timescales shorter than 2 days,
is incorrect (Mr´oz et al.
but one measurement, for MOA-ip-1,
2017).
The recent analysis of data from the Optical Grav-
itational Lensing Experiment (OGLE) (Udalski et al.
2015a) provides much stronger constraints on the abun-
dance of free-floating Jupiters. Mr´oz et al. (2017) ana-
lyzed a larger sample of over 2,600 microlensing events
discovered during the years 2010 -- 2015. They found that
Jupiter-mass lenses are at most an order of magnitude
less common than suggested by Sumi et al.
(with a
95% upper limit of 0.25 Jupiter-mass planets per main-
sequence star). They detected, however, a handful of
ultrashort-timescale microlensing events (with timescales
of less than 0.5 day), strongly suggesting the existence of
Earth-mass and super-Earth-mass free-floating planets.
Their light curves are not well covered with observations
from a single telescope, rendering the detection of the
finite-source effect difficult.
We conducted a pilot program of searching for ul-
trashort microlensing events in the 2016 observing sea-
son data. We supplemented OGLE observations with
data from the KMTNet survey, a network of longitude-
separated telescopes, which provided us with a better
coverage of short-timescale microlensing events.
Here, we present the discovery of an ultrashort-
timescale event OGLE-2016-BLG-1540 and report the
first measurement of the Einstein ring radius of a free-
floating planet candidate.
2. OBSERVATIONS
Microlensing event OGLE-2016-BLG-1540 was discov-
ered by the OGLE Early Warning System (Udalski 2003)
on 2016 August 6, at equatorial coordinates of R.A. =
18h00m47.s00 and Decl. = −28◦21(cid:48)35.(cid:48)(cid:48)2 (J2000.0), i.e.,
Galactic coordinates l = 2.186◦, b = −2.574◦. The sur-
vey uses a 1.3 m Warsaw Telescope at Las Campanas
Observatory in Chile (the Observatory is operated by
the Carnegie Institution for Science), equipped with a
1.4 deg2 mosaic CCD camera. The event was located in
field BLG512, which was observed with a cadence of 20
minutes.
The Korea Microlensing Telescope Network (KMTNet)
consists of three 1.6 m telescopes equipped with 4.0 deg2
cameras. The telescopes are located in CTIO (Chile),
SAAO (South Africa), and SSO (Australia), see Kim
et al. (2016) for details. The event was located in two
overlapping fields BLG03 and BLG43, monitored with a
cadence of 14 minutes. We omitted KMT SSO observa-
tions, because they did not cover the peak and did not
contribute to constraining the model. We also excluded
KMT SAAO data taken before August 3 or after August
16, because the baseline light curve showed systematic
variability connected with passages of the Moon near the
bulge fields.
All observations used in the modeling were taken in
the I band. Photometry was extracted using custom
implementations of the difference image analysis (Alard
& Lupton 1998): Wo´zniak (2000) (OGLE) and Albrow
et al. (2009) (KMTNet). The photometric uncertainties
were corrected using the standard procedures (Skowron
et al. 2016). We additionally reduced KMT CTIO V -
and I-band images using DoPhot (Schechter et al. 1993),
which allowed us to determine the source color.
3. LIGHT CURVE MODELING
A Neptune-mass Free-floating Planet Candidate Discovered by Microlensing Surveys
3
Fig. 1. -- Microlensing event OGLE-2016-BLG-1540 exhibits prominent finite-source effect, because the source is larger than the angular
Einstein ring. The light curve can be accurately described using the finite-source point-lens model (black solid line in the I-band, gray
dashed line in the V -band). I- and V -band models differ because of different limb-darkening profiles of the source star in two filters.
V -band data were not used in the modeling. All measurements were transformed to the OGLE magnitude scale.
The light curve of the event (Figure 1) can be accu-
rately described using the finite-source point-lens model.
The model has four parameters: the time and projected
separation of closest approach of the source to the lens
t0 and u0, the Einstein timescale tE, and the normalized
angular radius of the source ρ = θ∗/θE (θ∗ is the angular
radius of the source).
Two parameters, Fs and Fb, describe the source and
(unmagnified) blend fluxes, respectively. When we al-
lowed Fb to vary, we found that in the best-fit solution
the blend flux is negative (with the absolute value corre-
sponding to a 16 -- 17-mag star). Although such solutions
are mathematically possible, this negative blending is too
big to be due to normal fluctuations in the background.
The best-fit solution is only ∆χ2 = 5 better than the
solution with fixed Fb = 0, which can easily be due to
statistical noise, or possibly low-level systematics in the
data. Given the absence of evidence for blending and
the low prior probability for ambient superposed bright
source, our best estimate for the blended light is zero,
i.e., Fb = 0. The only way that the source flux enters
the characterization of the lens is via θ∗ (see Section 4.1).
To account for this, while we fix Fb = 0 in the fits, we
also add in quadrature 0.05 mag to the uncertainty in
centroiding the clump, when we compute our errors of
these quantities.
S(ϕ)/ ¯S = 1− Γ(1− 3
Two additional (wavelength-dependent) parameters Γ
and Λ may be used to describe the limb-darkening profile:
4√cos ϕ), where ϕ is
the angle between the normal to the stellar surface and
the line of sight (Yoo et al. 2004). The two-parameter
limb darkening law provides a more accurate description
of a brightness profile than a simple linear law (e.g., Al-
brow et al. 1999; Fields et al. 2003; Abe et al. 2003). We
2 cos ϕ)− Λ(1− 5
TABLE 1
Best-fitting model parameters
Parameter
t0 (HJD(cid:48))
tE (days)
u0
ρ
Is
fs
χ2/d.o.f.
Value
Uncertainty
7606.726
0.320
0.53
1.65
14.76
1.00
0.002
0.003
0.04
0.01
0.05
(fixed)
2160.1/2153
Note: HJD(cid:48)=HJD-2450000. fs = Fs/(Fs + Fb) is the blending
parameter.
used a fixed ΓI = 0.36 and ΛI = 0.34 which correspond
to the physical parameters of the source star (c.f., Sec-
tion 4.1). When we allowed Γ and Λ to vary, we found
Γ = 0.25±0.20 and Λ = 0.36±0.40, consistent at 1.5−2σ
level with the adopted values.
The finite-source magnifications were calculated by the
direct integration of formulae derived by Lee et al. (2009),
which remain valid in the low-magnification regime. The
uncertainties were estimated using the Markov Chain
Monte Carlo method. The best-fitting parameters and
their 1σ error bars are shown in Table 1.
We also considered models with terrestrial parallax
(Gould et al. 2009; Freeman et al. 2015). The microlens
parallax in the best-fitting solution was πE = 3200± 700,
but the χ2 improvement was modest (∆χ2 = 18).
The parallax signal came mostly from one observatory
(KMT CTIO) from one night and the OGLE data from
that night did not provide strong evidence for parallax.
Thus, the terrestrial parallax signal may be mimicked
by some low-level systematics in the data and cannot be
trusted.
14.214.314.414.514.614.714.8MagnitudeOGLEIKMTCTIOIKMTCTIOVKMTSAAOI76037604760576067607760876097610HJD-2450000−0.04−0.020.000.020.04Residual4
Mr´oz et al.
4. PHYSICAL PARAMETERS
4.1. Source star
The event was observed in the V -band by the
KMT CTIO on the peak night (Figure 1), which al-
lowed us to measure the color of the source and hence
the Einstein ring radius θE (Yoo et al. 2004). Because
the finite-source effect is prominent and the event may
no longer be achromatic, we have not used the model-
independent regression to estimate the source color. In-
stead, we calculated the color for each link of the MCMC
chain. The procedure of model fitting, calculating the
source color and the limb-darkening coefficients was iter-
ated, until the color measurement converged. We com-
pared the location of the source and red clump centroid
in the instrumental color -- magnitude diagram (CMD) in
× 2(cid:48) region around the event for KMT CTIO. We
a 2(cid:48)
found that the source is ∆(V − I) = 0.61 ± 0.02 redder
and ∆I = −0.85 ± 0.09 brighter than the red clump.
Assuming the intrinsic color of (V − I)RC,0 = 1.06 of
red clump stars (Bensby et al. 2011) and their mean de-
reddened brightness in this direction of IRC,0 = 14.36
(Nataf et al. 2013), we calculated the intrinsic brightness
IS,0 = 13.51 ± 0.09 and color (V − I)S,0 = 1.67 ± 0.02 of
the source. The OGLE-IV CMD, for a larger region of
× 4(cid:48), is shown in Figure 2.
4(cid:48)
We then found (V − K)S,0 = 3.67 ± 0.03 from color --
color relations from Bessell & Brett (1988) and estimated
the angular radius of the source star θ∗ = 15.1 ± 0.8 µas
from color-surface brightness relation for giants (Kervella
et al. 2004). The latter estimate allowed us to measure
the angular Einstein radius
θE = θ∗/ρ = 9.2 ± 0.5 µas
and the relative lens-source proper motion
µrel = θE/tE = 10.5 ± 0.6 mas yr−1.
We can also estimate the effective temperature of the
source of Teff = 3900±200 K using the color -- temperature
relations of Houdashelt et al. (2000) and Ram´ırez &
Mel´endez (2005). The corresponding limb-darkening co-
efficients (Claret & Bloemen 2011) are:
ΓI =0.36 ΛI =0.34
ΓV =0.94 ΛV =−0.21
We used ATLAS models and assumed a solar metallic-
ity, microturbulent velocity of 2 km/s and surface grav-
ity of log g = 2.0. Limb-darkening coefficients (c, d) from
Claret & Bloemen (2011) were transformed to (Γ, Λ) us-
ing formulae derived by Fields et al. (2003).
4.2. Source proper motion
As the source star is relatively bright and the contri-
bution of the lens to the total light is negligible, it is
possible to measure the absolute proper motion of the
source. We chose a subset of 363 best-seeing (0.7 − 1(cid:48)(cid:48))
and low-background images (out of 11,276 total epochs of
this field) spanning 2010 -- 2017 taken with the 24th CCD
detector of the OGLE-IV camera. We used the CMD to
identify 3818 candidate red clump stars, which served as
anchors for the coordinates transformations between the
CCD frames. This allowed us to measure proper motions
with respect to the mean motion of the Galactic bulge.
Fig. 2. -- The OGLE-IV color -- magnitude diagram for stars in
a 4(cid:48) × 4(cid:48) region around OGLE-2016-BLG-1540. The source (blue
square) is located in a relatively lowly populated region of the
diagram. Blue and red areas mark stars used for the proper motion
measurements (see Section 4.2).
We used the DoPhot PSF photometry package
(Schechter et al. 1993) to measure positions of all stars in
all 363 images. Then, we calculated the third-order poly-
nomial coordinate transformations between each frame
and the first frame by minimizing the scatter for the
anchor red clump stars. The proper motions were fit-
ted with the least-squares method (with outlier rejec-
tion). The formal uncertainties of the fit were typi-
cally 0.2 − 0.3 mas yr−1. However, the comparison with
proper motion measurements based on the OGLE-III
data (2001 -- 2009) showed discrepancies larger than the
pure statistical error. We decided to employ 0.5 and
0.7 mas yr−1 for N and E directions as our measure of
uncertainty; hence, the proper motion of the source is
µS = (µN, µE) = (−5.6 ± 0.5,−3.0 ± 0.7) mas yr−1 with
respect to the Galactic bulge (see Figure 3).
As both the position of the source star on the CMD is
uncommon (the star is located below the red giant branch
and redwards of red clump giants) and its proper motion
is counter to the Galactic rotation, we consider whether
this evidence indicates that the source belongs to the far
disk population. First, we investigated the CMD position
of the source. We identified about 45 stars with similar
CMD positions3 and measured their kinematics (their
proper motions are marked with gray crosses in Figure
3), finding that these are consistent with all other red
giant (i.e., bulge) stars in the field4. Thus, the unusual
3 The analyzed region is marked with a blue rectangle in Figure
2.
4 We compared both distributions of proper motions using
the two-sample Anderson -- Darling test and found p-values of 0.21
(for µl component) and 0.65 (µb).
Similarly, the two-sample
Kolmogorov -- Smirnov test yields p-values of 0.29 and 0.34, respec-
tively. Therefore, there is no evidence that these distributions are
different.
0.51.01.52.02.53.03.5V−I(OGLE-IV)121314151617181920I(OGLE-IV)redclumpsourceA Neptune-mass Free-floating Planet Candidate Discovered by Microlensing Surveys
5
tion is slightly weaker for larger mass ratios (because the
host event is shorter). For q = 10−3, corresponding to
the 0.5 M(cid:12) host in the bulge (πrel = 0.01 mas), we found
s > 4.8.
We note that the presence of a putative host may also
be revealed by perturbations to the point-lens light curve
due to the planetary caustic caused by the central star
(Han & Kang 2003; Han et al. 2005). The angular size of
the planetary caustic (relative to the Einstein radius of
the planet) is 4/s2 ≤ 0.16 for s ≥ 5. Because the source
star is ∼ 10 times larger, the signatures of the caustic
are washed out by the finite-source effect.
5. DISCUSSION
Current microlensing surveys are capable of detecting
free-floating planets down to Earth-mass objects. To
this day, however, all reported free-floating planet candi-
dates were based on the very short timescale of an event
(tE (cid:46) 2 days) and lacked direct measurements of the an-
gular Einstein ring size (Sumi et al. 2011; Mr´oz et al.
2017). OGLE-2016-BLG-1540 is the first case for which
we procured such a measurement, owing to the fortu-
itous fact that the source was a giant. If the source were
a dwarf (with at least ten times smaller angular radius),
as in the case of ultrashort candidate events detected by
Mr´oz et al. (2017), the finite-source effect would be signif-
icantly weaker. We simulated the OGLE light curve and
found that the finite-source model would be preferred
only by ∆χ2 = 1.6 over the point-lens model.
The short timescale of the event can be explained in
part by the unusual kinematics of the system (see Fig-
ure 3). The source is moving at µS = 6.4 mas yr−1
in the direction opposite to the Galactic rotation and
the relative lens-source proper motion is large (µrel =
10.5 ± 0.6 mas yr−1). One possible explanation is that
the source is located behind the Galactic center in the far
disk, in which case we expect the proper motion direc-
tion to be opposite compared to closer stars. To test the
"far disk" hypothesis, we have studied proper motions of
stars located near the source in the CMD (Section 4.2).
These stars follow exactly the same distribution of proper
motions as the bulge stars.
It appears that, although
the source proper motion has an unusual direction, the
source belongs to the bulge population.
The large lens-source proper motion indicates that the
lens is moving in the opposite direction than the source
(along the Galactic rotation) at µL (cid:38) 5 mas yr−1 relative
to red clump stars. The gray dashed circle in Figure 3
marks the relative proper motion of µrel = 10.5 mas yr−1
with respect to the source. As the lens should be lo-
cated on this circle, it likely belongs to the Galactic disk
population. Only 15% of bulge stars (58% of the disk
stars) are located outside the dashed circle in Figure 3,
i.e., their proper motions with respect to the source star
are higher than µrel.
Because the distance to the lens, and so the relative
parallax πrel, is unknown, we cannot uniquely measure
the lens mass (eq. (1)):
M =
θ2
E
κπrel
= 35 M⊕
0.1 mas
πrel
If the lens is located in the disk (πrel ≈ 0.1 mas), it
should be a Neptune-mass planet. However, we cannot
Fig. 3. -- Proper motions of stars in the OGLE-2016-BLG-1540
field (9(cid:48)×18(cid:48)). Orange contours mark proper motions of red clump
stars (bulge population), black contours mark main-sequence stars
(which represent the Galactic disk population). The dashed red
line shows the direction of increasing Galactic longitude and the
proper motion of the source star is marked with a blue dot. The
gray dashed circle corresponds to the relative proper motion of 10.5
mas yr−1 with respect to the source. As the lens should be located
on this circle, it likely belongs to the Galactic disk population.
Gray crosses mark stars located near the source in the CMD; they
follow the bulge distribution.
position of the source star on the CMD cannot be taken
as evidence for belonging to some other population.
4.3. Constraints on the host star
Additional features in the event light curve would have
been detected if the trajectory of the source were for-
tunate enough to pass near a putative host star. Be-
cause the light curve does not exhibit any signatures of
the host star, we can only provide a lower limit on the
planet-host separation, using a variation of the method
proposed by Gaudi & Sackett (2000). We simulated ar-
tificial OGLE light curves (spanning from 2010 March
4, through 2017 October 10) for a given binary model,
defined by three additional parameters as compared to
the single-lens case: mass ratio q, star-planet separation
s (expressed in Einstein radii), and angle α between the
source trajectory and binary axis. (The remaining pa-
rameters were calculated based on the best-fitting single-
lens model from Table 1). We fitted binary and single
lensing models to the artificial data and calculated the
difference ∆χ2 = χ2
single− χ2
which corresponds to the Einstein radius of the host of
θE,host = θE/√q = 0.65 mas.
If the lens is located in
the Galactic disk (πrel = 0.1 mas), the corresponding
host mass is Mhost = 0.5 M(cid:12). For each value of s, we
simulated 180 light curves with uniformly distributed
α ∈ [0, 2π], and calculated the probability of detecting
the host star as the fraction of light curves which fulfill
∆χ2 > ∆χ2
thresh = 225. This probability drops below
90% when s > 5.1. The lower limit on the host separa-
binary. We used q = 2× 10−4,
1050510µE[masyr−1]1050510µN[masyr−1]Red clump stars (3722)Main Sequence stars (1251)source6
Mr´oz et al.
rule out that the lens belongs to the bulge population
(πrel ≈ 0.02 mas), in which case it should be a Saturn-
mass object. The geometry required for the lens to be
a brown dwarf (πrel (cid:46) 10−3 mas; i.e., DS − DL (cid:46) 60 pc)
would require significant fine-tuning. We note that the
event occurred inside the K2 C9 superstamp (Henderson
et al. 2016), but unfortunately the parallax measurement
was impossible, because the K2 C9 campaign finished a
few weeks before the event.
Microlensing alone cannot distinguish between wide-
orbit and unbound planets and, in principle, the lens
may be located at a wide orbit, like Uranus or Neptune.
Our lower limit for the planet-host separation is 5.1 Ein-
stein radii, which corresponds to the projected physical
separation of 15 au at πrel = 0.1 mas. Owing to the rel-
atively large lens-source proper motion, any stellar com-
panions to the lens can be detected in the future, when
the lens and source separate. However, the brightness of
the source will make detection of the putative host light
difficult.
The characterization of this event would have been im-
possible without nearly continuous observations from the
OGLE and KMTNet surveys. This event also shows the
importance of securing the color information for short-
timescale events and anomalies (see discussion in Hwang
et al. 2017). Although we could not have measured the
precise mass of the lens, such measurements will be pos-
sible in the future with the Euclid (Penny et al. 2013)
and WFIRST (Spergel et al. 2015) satellites, but will
require simultaneous ground- and space-based observa-
tions (Gould & Yee 2013; Yee 2013; Zhu & Gould 2016).
Current ground-based experiments are already sensitive
to ultrashort microlensing events, but a bigger sample is
needed to fully understand their origin.
ACKNOWLEDGEMENTS
We thank the anonymous referee for constructive com-
ments. The OGLE project has received funding from
the National Science Center, Poland, grant MAESTRO
2014/14/A/ST9/00121 to A.U. Work by W.Z., Y.K.J.,
and A.G. were supported by AST-1516842 from the US
NSF. W.Z., I.G.S., and A.G. were supported by JPL
grant 1500811. This research has made use of the KMT-
Net system operated by the Korea Astronomy and Space
Science Institute (KASI) and the data were obtained at
three host sites of CTIO in Chile, SAAO in South Africa,
and SSO in Australia. Work by K.H.H. was support by
KASI grant 2017-1-830-03. Work by C.H. was supported
by the grant (2017R1A4A101517) of National Research
Foundation of Korea.
REFERENCES
Abe, F., Bennett, D. P., Bond, I. A., et al. 2003, A&A, 411, L493
Alard, C., & Lupton, R. H. 1998, ApJ, 503, 325
Albrow, M. D., Beaulieu, J.-P., Caldwell, J. A. R., et al. 1999,
ApJ, 522, 1011
Albrow, M. D., Horne, K., Bramich, D. M., et al. 2009, MNRAS,
397, 2099
Alcock, C., Allen, W. H., Allsman, R. A., et al. 1997, ApJ, 491,
436
Barclay, T., Quintana, E. V., Raymond, S. N., & Penny, M. T.
2017, ApJ, 841, 86
Bennett, D. P., & Rhie, S. H. 1996, ApJ, 472, 660
Bensby, T., Ad´en, D., Mel´endez, J., et al. 2011, A&A, 533, A134
Bessell, M. S., & Brett, J. M. 1988, PASP, 100, 1134
Clanton, C., & Gaudi, B. S. 2017, ApJ, 834, 46
Claret, A., & Bloemen, S. 2011, A&A, 529, A75
Dong, S., Udalski, A., Gould, A., et al. 2007, ApJ, 664, 862
Fields, D. L., Albrow, M. D., An, J., et al. 2003, ApJ, 596, 1305
Freeman, M., Philpott, L. C., Abe, F., et al. 2015, ApJ, 799, 181
Gaudi, B. S., & Sackett, P. D. 2000, ApJ, 528, 56
Gould, A. 1992, ApJ, 392, 442
-- . 2016, Journal of Korean Astronomical Society, 49, 123
Gould, A., & Yee, J. C. 2013, ApJ, 764, 107
Gould, A., Udalski, A., Monard, B., et al. 2009, ApJ, 698, L147
Han, C., Gaudi, B. S., An, J. H., & Gould, A. 2005, ApJ, 618, 962
Han, C., & Kang, Y. W. 2003, ApJ, 596, 1320
Henderson, C. B., & Shvartzvald, Y. 2016, AJ, 152, 96
Henderson, C. B., Poleski, R., Penny, M., et al. 2016, PASP, 128,
124401
Houdashelt, M. L., Bell, R. A., & Sweigart, A. V. 2000, AJ, 119,
1448
Hwang, K.-H., Udalski, A., Bond, I. A., et al. 2017, ArXiv
e-prints, arXiv:1711.09651
Kervella, P., Bersier, D., Mourard, D., et al. 2004, A&A, 428, 587
Kim, S.-L., Lee, C.-U., Park, B.-G., et al. 2016, Journal of Korean
Astronomical Society, 49, 37
Lee, C.-H., Riffeser, A., Seitz, S., & Bender, R. 2009, ApJ, 695,
200
Ma, S., Mao, S., Ida, S., Zhu, W., & Lin, D. N. C. 2016, MNRAS,
461, L107
Mao, S., Di Stefano, R., Alcock, C., et al. 1994, in Bulletin of the
American Astronomical Society, Vol. 26, American
Astronomical Society Meeting Abstracts, 1336
Mr´oz, P., Udalski, A., Skowron, J., et al. 2017, Nature, 548, 183
Nataf, D. M., Gould, A., Fouqu´e, P., et al. 2013, ApJ, 769, 88
Nemiroff, R. J., & Wickramasinghe, W. A. D. T. 1994, ApJ, 424,
L21
Paczy´nski, B. 1986, ApJ, 304, 1
Pena Ram´ırez, K., B´ejar, V. J. S., Zapatero Osorio, M. R.,
Petr-Gotzens, M. G., & Mart´ın, E. L. 2012, ApJ, 754, 30
Penny, M. T., Rattenbury, N. J., Gaudi, B. S., & Kerins, E. 2017,
AJ, 153, 161
Penny, M. T., Kerins, E., Rattenbury, N., et al. 2013, MNRAS,
434, 2
Pfyffer, S., Alibert, Y., Benz, W., & Swoboda, D. 2015, A&A,
579, A37
Ram´ırez, I., & Mel´endez, J. 2005, ApJ, 626, 465
Refsdal, S. 1966, MNRAS, 134, 315
Schechter, P. L., Mateo, M., & Saha, A. 1993, PASP, 105, 1342
Scholz, A., Jayawardhana, R., Muzic, K., et al. 2012, ApJ, 756, 24
Skowron, J., Udalski, A., Koz(cid:32)lowski, S., et al. 2016, Acta Astron.,
66, 1
Spergel, D., Gehrels, N., Baltay, C., et al. 2015, ArXiv e-prints,
arXiv:1503.03757
Sumi, T., Kamiya, K., Bennett, D. P., et al. 2011, Nature, 473,
349
Udalski, A. 2003, Acta Astron., 53, 291
Udalski, A., Szymanski, M., Mao, S., et al. 1994, ApJ, 436, L103
Udalski, A., Szyma´nski, M. K., & Szyma´nski, G. 2015a, Acta
Astron., 65, 1
Udalski, A., Yee, J. C., Gould, A., et al. 2015b, ApJ, 799, 237
Veras, D., & Raymond, S. N. 2012, MNRAS, 421, L117
Wo´zniak, P. R. 2000, Acta Astron., 50, 421
Yee, J. C. 2013, ApJ, 770, L31
Yoo, J., DePoy, D. L., Gal-Yam, A., et al. 2004, ApJ, 603, 139
Zhu, W., & Gould, A. 2016, Journal of Korean Astronomical
Society, 49, 93
Zhu, W., Huang, C. X., Udalski, A., et al. 2017, PASP, 129,
104501
|
1004.1662 | 2 | 1004 | 2010-04-13T22:51:20 | Axisymmetric modes in vertically stratified self-gravitating discs | [
"astro-ph.EP",
"astro-ph.GA",
"astro-ph.SR"
] | We perform linear analysis of axisymmetric vertical normal modes in stratified compressible self-gravitating polytropic discs in the shearing box approximation. We study specific dynamics for subadiabatic, adiabatic and superadiabatic vertical stratifications. In the absence of self-gravity, four well-known principal modes can be identified in a stratified disc: acoustic p-, surface gravity f-, buoyancy g- and inertial r-modes. After characterizing modes in the non-self-gravitating case, we include self-gravity and investigate how it modifies the properties of these modes. We find that self-gravity, to a certain degree, reduces their frequencies and changes the structure of the dispersion curves and eigenfunctions at radial wavelengths comparable to the disc height. Its influence on the basic branch of the r-mode, in the case of subadiabatic and adiabatic stratifications, and on the basic branch of the g-mode, in the case of superadiabatic stratification (which in addition exhibits convective instability), does appear to be strongest. Reducing the three-dimensional Toomre's parameter Q_{3D} results in the latter modes becoming unstable due to self-gravity, so that they determine the onset criterion and nature of gravitational instability of a stratified disc. By contrast, the p-, f- and convectively stable g-modes, although their corresponding \omega^2 are reduced by self-gravity, never become unstable however small the value of Q_{3D}. This is a consequence of the three-dimensionality of the disc. The eigenfunctions corresponding to the gravitationally unstable modes are intrinsically 3D. We also contrast the more exact instability criterion based on our 3D model with that of density waves in 2D (razor-thin) discs. Based on these findings, we comment on the origin of surface distortions seen in numerical simulations of self-gravitating discs. | astro-ph.EP | astro-ph | Mon. Not. R. Astron. Soc. 000, 1 -- ?? (2010)
Printed 29 October 2018
(MN LATEX style file v2.2)
Axisymmetric modes in vertically stratified self-gravitating
discs
G. R. Mamatsashvili1,2⋆ and W. K. M. Rice 1
1 SUPA, Institute for Astronomy, University of Edinburgh, Blackford Hill, Edinburgh EH9 3HJ, Scotland
2 Georgian National Astrophysical Observatory, Ilia State University, 2a Kazbegi Ave., Tbilisi 0160, Georgia
Accepted 2010 April 09. Received 2010 April 08; in original form 2010 February 09
ABSTRACT
We carry out a linear analysis of the vertical normal modes of axisymmetric pertur-
bations in stratified, compressible, self-gravitating gaseous discs in the shearing box
approximation. An unperturbed disc has a polytropic vertical structure that allows
us to study specific dynamics for subadiabatic, adiabatic and superadiabatic vertical
stratifications, by simply varying the polytropic index. In the absence of self-gravity,
four well-known principal modes can be identified in a stratified disc: acoustic p-modes,
surface gravity f-modes, buoyancy g-modes and inertial r-modes. After classifying and
characterizing modes in the non-self-gravitating case, we include self-gravity in the
perturbation equations and in the equilibrium and investigate how it modifies the
properties of these four modes. We find that self-gravity, to a certain degree, reduces
their frequencies and changes the structure of the dispersion curves and eigenfunc-
tions at radial wavelengths comparable to the disc height. Its influence on the basic
branch of the r-mode, in the case of subadiabatic and adiabatic stratifications, and
on the basic branch of the g-mode, in the case of superadiabatic stratification (which
in addition exhibits convective instability), does appear to be strongest. Reducing the
three-dimensional Toomre's parameter Q3D results in the latter modes becoming un-
stable due to self-gravity, so that they determine the onset criterion and nature of the
gravitational instability of a vertically stratified disc. By contrast, the p-, f- and con-
vectively stable g-modes, although their corresponding ω2 are reduced by self-gravity,
never become unstable however small the value of Q3D. This is a consequence of the
three-dimensionality of the disc. The eigenfunctions corresponding to the gravitation-
ally unstable modes are intrinsically three-dimensional. We also contrast the more
exact instability criterion based on our three-dimensional model with that of density
waves in two-dimensional (razor-thin) discs. Based on these findings, we comment on
the origin of surface distortions seen in numerical simulations of self-gravitating discs.
Key words: accretion, accretion discs -- gravitation -- hydrodynamics -- instabilities
-- convection -- (stars:)planetary systems: protoplanetary discs -- turbulence
0
1
0
2
r
p
A
3
1
.
]
P
E
h
p
-
o
r
t
s
a
[
2
v
2
6
6
1
.
4
0
0
1
:
v
i
X
r
a
1
INTRODUCTION
Self-gravity plays an important role in a variety of as-
trophysical systems. It is a main agent determining the
dynamical evolution of star clusters, galaxies, various
types of accretion discs, etc. Particularly in protoplanetary
discs, that are the central subject of our study, self-gravity
provides one of the main source of outward angular mo-
mentum transport through the excitation of density waves
(Papaloizou & Savonije
1991; Laughlin & Bodenheimer
1994; Lodato & Rice 2004, 2005) and is able to cause frag-
mentation of a disc into bound clumps, or planets, under
⋆ E-mail: [email protected]
certain conditions (Gammie 2001; Rice et al. 2003, 2005;
Boss 2004; Mayer et al. 2007; Rafikov 2007). Starting with
the seminal paper by Toomre (1964), there have been a large
number of studies of the stability of self-gravitating gaseous
discs, both in the linear (e.g., Goldreich & Lynden-Bell
1965a,b; Goldreich & Tremaine 1978; Adams et al. 1989;
Bertin et al. 1989; Papaloizou & Lin 1989; Laughlin et al.
rele-
1997) and non-linear
radiation
vant physical
factors
transport) with up-to-date numerical
(e.g.,
Papaloizou & Savonije
Laughlin & Bodenheimer
1994; Boss 1998; Pickett et al. 2000, 2003; Gammie 2001;
Boss 2003; Johnson & Gammie 2003; Rice et al. 2003, 2005;
(e.g., heating, cooling,
techniques
regimes,
including other
1991;
2 G. R. Mamatsashvili and W. K. M. Rice
Boss 2004; Mej´ıa et al. 2005; Boley et al. 2006; Mayer et al.
2007; Stamatellos & Whitworth 2009).
Linear stability analysis in a vast majority of cases is
restricted, for simplicity, to razor-thin, or two-dimensional
(2D) discs that are obtained by vertically averaging all quan-
tities. In other words, perturbations are assumed to have
large horizontal scales compared with the disc thickness. In
this limit, the well-known Toomre's parameter
Q2D =
csΩ
πGΣ
controls the stability of self-gravitating discs (Toomre 1964).
In this case, a density wave, which is the only mode in
a 2D disc, is influenced by self-gravity and thus can be-
come unstable, as the local dispersion relation for the
latter clearly demonstrates (Goldreich & Tremaine 1978;
Binney & Tremaine 1987; Bertin et al. 1989).
Stability analysis in a more realistic case of self-
gravitating three-dimensional (3D) discs is more compli-
cated. The disc is vertically stratified due to both its own
self-gravity and the vertical component of the gravity of a
central object. Depending on the nature of the stratification,
there exists a whole new set of various vertical modes in the
system (see below), some of which can become unstable due
to self-gravity on horizontal length scales comparable to the
disc thickness. In this situation, the vertical variation of per-
turbations is important and for a correct characterization of
the gravitational instability it is necessary to introduce an-
other parameter not involving height-dependent variables,
such as the sound speed in Toomre's parameter. Further-
more, not all types of stratification permit two-dimensional
modes, that is, modes with no vertical motions commonly
occurring in the 2D treatment. For example, in non-self-
gravitating discs with polytropic vertical structure, there
are no 2D modes (Lin et al. 1990; Lubow & Pringle 1993b;
Korycansky & Pringle 1995) implying that the dynamics
does not always reduce to that of the 2D case. Therefore,
a more accurate stability analysis of self-gravitating discs
should necessarily be three-dimensional.
Obviously, before studying the gravitational instability
of stratified discs, one must first classify and characterize
vertical normal modes of perturbations in the simplified case
of no self-gravity. Analysis of the modal structure of strati-
fied, polytropic, compressible, non-self-gravitating discs has
been done in several papers: Ruden et al. (1988, hereafter
RPL), Korycansky & Pringle (1995, hereafter KP), Ogilvie
(1998). In convectively stable discs, i.e., with subadiabatic
vertical stratification, four principal types of vertical modes
can be distinguished. These modes are: acoustic p-modes,
surface gravity f-modes, buoyancy g-modes and inertial
r-modes. The modes are named after their correspond-
ing restoring forces, which can be well identified for each
mode at horizontal wavelengths smaller than the disc height
and are provided by one of the following: compressibil-
ity/pressure, displacements of free surfaces of a disc, buoy-
ancy due to vertical stratification and inertial forces due to
disc rotation, respectively, for the p-, f-, g- and r-modes. In
the case of superadiabatic stratification, the r- and g-modes
merge and appear as a single mode, which becomes con-
vectively unstable for horizontal wavenumbers larger than
a certain value (RPL); the p- and f-modes remain qualita-
tively unchanged. For neutral/adiabatic stratification, buoy-
ancy is absent and the g-mode disappears. The main pur-
pose of this paper is to investigate how self-gravity modifies
the frequencies and the structure of the eigenfunctions of
these modes, which mode acquires the largest positive growth
rate due to self-gravity and, therefore, determines the on-
set criterion and nature of the gravitational instability of a
stratified disc. So, the mode dynamics in the 3D case can ap-
pear more complex than that in the 2D one, where only the
density wave mode can be subject to gravitational instabil-
ity. Previously, Goldreich & Lynden-Bell (1965a, hereafter
GLB) considered gravitational instability in a uniformly ro-
tating gaseous slab with an adiabatic vertical stratification,
thereby leaving out all modes associated with buoyancy.
Other studies also considered the gravitational instability
of 3D galactic discs, however, the analysis was essentially
2D, finite-thickness effects were only taken into account by
means of various reduction factors in 2D dispersion rela-
tions (Shu 1968; Romeo 1992, 1994). In all these studies, as
in GLB, the main focus was on finding the criterion for the
onset of gravitational instability, so that a full analysis of
various types of vertical normal modes existing in stratified
self-gravitating discs was not carried out. Actually, we gen-
eralize the study of GLB to subadiabatic and superadiabatic
stratifications having different modal structure.
Another motivation for our study is that the f-mode
is thought to play an important dynamical role in self-
gravitating discs. The non-linear behaviour of 3D perturba-
tions involving large surface distortions, as seen in numer-
ical simulations, has been attributed to the surface grav-
ity f-mode (Pickett et al. 2000). However, this was done
without analysing the behaviour of other vertical modes
under self-gravity. It was shown that the f-mode leads to
a large energy dissipation in the vicinity of the disc sur-
face, which may facilitate disc cooling, because the energy
is deposited at smaller optical depth where it can be radi-
ated away more quickly (see e.g., Johnson & Gammie 2003;
Boley et al. 2006). Later it was realized that in fact the non-
linear vertical motions in self-gravitating discs can be much
more complex than just the f-mode and can have a shock
character (shock bores, Boley & Durisen 2006). Thus, in the
3D case, the dynamics of self-gravitating discs is much richer
and diverse than that of idealized 2D ones and requires fur-
ther study. To fully understand the origin of such three-
dimensional effects and what type of instability they are as-
sociated with, one should start with a rigorous linear study
of the characteristic properties of all the types of vertical
normal modes mentioned above, not only the f-mode, in the
presence of self-gravity. The present work is just a first step
in this direction.
Numerical simulations of self-gravitating discs are of-
ten in the context of global discs (e.g., Pickett et al. 2000,
2003; Rice et al. 2003, 2005; Lodato & Rice 2004, 2005;
Boley et al. 2006; Boley 2009) and, therefore, are not al-
ways able to well resolve vertical motions, which, as shown
in the present study, inevitably arise during the develop-
ment of the gravitational instability associated with intrin-
sically three-dimensional modes. So, these simulations may
not quite accurately capture all the subtleties of the gravita-
tional instability in 3D discs. In this connection, we should
mention the work by Nelson (2006) that extensively dis-
cusses the issue of vertical resolution and its importance
in the outcome of the gravitational instability in numerical
Axisymmetric modes in vertically stratified self-gravitating discs
3
simulations of self-gravitation discs. Resolving and analysing
vertical motions are also crucial for properly understanding
cooling processes in discs and, particularly, whether convec-
tion is able to provide sufficiently effective cooling for disc
fragmentation to occur, which is still a matter for debate
in the literature (Boss 2004; Mayer et al. 2007; Boley et al.
2006, 2007; Rafikov 2007). In addition, these studies, for sim-
plicity, use the criterion for gravitational instability based
on the two-dimensional Toomre's parameter, which, as we
will demonstrate, cannot always be uniquely mapped into an
analogous three-dimensional stability parameter and give a
precise criterion for the onset of gravitational instability.
In this paper,
following other works
in a simi-
lar vein: Lubow & Pringle (1993b, hereafter LP), KP,
Lubow & Ogilvie (1998, hereafter LO98), we adopt the
shearing box approximation and consider the linear dynam-
ics of vertical normal modes of perturbations in a compress-
ible, stratified, self-gravitating gaseous disc with Keplerian
rotation. In the unperturbed disc, pressure and density are
related by a polytropic law, which is a reasonably good ap-
proximation for optically thick discs (see e.g., LO98). This
allows us to consider the specific features of the dynam-
ics for subadiabatic, adiabatic and superadiabatic vertical
stratifications by simply varying the polytropic index. As a
first step towards understanding the effects of self-gravity
on the perturbation modes, we restrict ourselves to axisym-
metric perturbations only. The linear results obtained here
will form the basis for studying the non-linear development
of gravitational instability in the local shearing box approx-
imation that allows much higher numerical resolution than
global disc models can afford.
The paper is organized as follows. The physical model
and basic equations are introduced in Section 2. The clas-
sification of vertical modes in the absence of self-gravity is
performed in Section 3. Effects of self-gravity on all normal
modes are analysed in Section 4. In Section 5, we focus on
the properties of gravitational instability in 3D. Compari-
son between the criteria of gravitational instability in 3D
and 2D is made in Section 6. Summary and discussions are
given in Section 7.
2 PHYSICAL MODEL AND EQUATIONS
in gaseous
self-gravitating discs,
three-dimensional
In order to study the dynamics of
modes
following LP,
KP, LO98, we adopt a local shearing box approxima-
tion (Goldreich & Lynden-Bell 1965b). In the shearing box
model, the disc dynamics is studied in a local Cartesian ref-
erence frame rotating with the angular velocity of disc ro-
tation at some fiducial radius from the central star, so that
curvature effects due to cylindrical geometry of the disc are
ignored. In this coordinate frame, the unperturbed differ-
ential rotation of the disc manifests itself as a parallel az-
imuthal flow with a constant velocity shear in the radial di-
rection. A Coriolis force is included to take into account the
effects of coordinate frame rotation. The vertical component
of the gravity force of the central object is also present. As
a result, we can write down the three-dimensional shearing
box equations
du
dt
+ 2Ωz × u + ∇(cid:16)−qΩ2x2 +
1
2
Ω2z2 + ψ(cid:17) +
1
ρ∇p = 0, (1)
dρ
dt
+ ρ∇ · u = 0
and the equation of conservation of entropy
d
dt (cid:18) p
ργ(cid:19) = 0.
(2)
(3)
This set of equations is supplemented with Poisson's equa-
tion to take care of disc self-gravity
(cid:18) ∂ 2
∂x2 +
∂ 2
∂y2 +
∂ 2
∂z2(cid:19) ψ = 4πGρ.
(4)
Here u = (ux, uy, uz) is the velocity in the local frame, p, ρ, ψ
are, respectively, the pressure, density and the gravitational
potential of the disc gas. Ω is the angular velocity of the local
reference frame rotation as a whole, x, y, z are, respectively,
the radial, azimuthal and vertical coordinates, z is the unit
vector along the vertical direction and d/dt = ∂/∂t + u · ∇
is the total time derivative. The shear parameter is q =
1.5 for the Keplerian rotation considered in this paper. The
adiabatic index, or the ratio of specific heats, γ = 1.4 as
typical of a disc composed of molecular hydrogen; we adopt
this value throughout the paper.
2.1 The equilibrium disc model
Equations (1-4) have an equilibrium solution that is station-
ary and axisymmetric. In this unperturbed state, the veloc-
ity field represents, as noted above, a parallel azimuthal flow,
u0, with a constant radial shear q:
ux0 = uz0 = 0, uy0 = −qΩx.
1
ρ0
dp0
dz
= Ω2z +
In the shearing box, the equilibrium density ρ0, pressure p0
and gravitational potential ψ0 depend only on the vertical
z−coordinate and satisfy the hydrostatic relation
g0 ≡ −
d2ψ0
dz2 = 4πGρ0.
As in LP, KP and RPL, pressure and density in the unper-
turbed disc are related by a polytropic relationship of the
form
dψ0
dz
(6)
(5)
,
p0 = Kρ1+1/s
0
,
(7)
where K is the polytropic constant and s > 0 is the poly-
tropic index. The Brunt-Vaisala frequency squared is defined
as
g0
c2
s
,
(8)
dρ0
0
c2
s
dp0
dz −
ρ0 (cid:16) 1
s + 1 − 1(cid:17) g2
dz (cid:17) = (cid:16) γs
N 2
0 ≡
where c2
s = γp0/ρ0 is the adiabatic sound speed squared.
If 1 + 1/s < γ (subadiabatic thermal stratification), then
N 2
0 > 0 all along the height and the equilibrium vertical
structure of the disc is convectively stable. In the oppo-
site case 1 + 1/s > γ (superadiabatic thermal stratifica-
tion), N 2
0 < 0 everywhere and this corresponds to a convec-
tively unstable equilibrium. For 1 + 1/s = γ (adiabatic ther-
mal stratification), N 2
0 = 0 and all the motions/modes due
to buoyancy disappear. Later we will consider these three
regimes separately.
4 G. R. Mamatsashvili and W. K. M. Rice
To determine the vertical structure, we need to solve
equations (5-6) subject to the boundary condition that the
pressure vanish at the free surface of the disc. Because we
have a polytropic model, it is convenient to work with the
pseudo-enthalpy
w0 = (s + 1)Kρ1/s
0
.
The disc structure is also symmetric with respect to the mid-
plane, z = 0, and, as a consequence, it follows from equa-
tions (5-7) that the derivative of w0 at the midplane vanishes
(i.e., dw0/dz(0) = 0). Because of this reflection symmetry,
we consider only the upper half of the disc, z > 0. At the
surface of the disc w0 = 0, similar to the pressure. This al-
lows us to determine the equilibrium vertical structure of the
disc. The non-dimensional variables being used throughout
the paper are:
1
0.8
0.6
0.4
0.2
0
0
1
0.8
0.6
0.4
0.2
ρ
0
ρ
0
0
0
0.5
s=5
1
2
3
s=1.7
z
1
z
1
0.8
0.6
0.4
0.2
0
0
1
0.8
0.6
0.4
0.2
2
c
s
2
c
s
s=5
1
2
3
z
1
z
s=1.7
1.5
2
1.5
2
0
0
0.5
x →
ρ0 →
xΩ
csm
ρ0
ρm
,
y →
, w0 →
,
yΩ
csm
w0
wm
z →
,
cs →
,
zΩ
csm
cs
csm
,
where ρm ≡ ρ0(0), wm ≡ w0(0), csm ≡ cs(0) are the mid-
plane values of the equilibrium density, pseudo-enthalpy and
sound speed. We define the three-dimensional analogue of
Toomre's parameter as
Q3D =
Ω2
4πGρm
,
which is a measure of disc self-gravity (from now on until
Section 6, we will use Q3D without subscript everywhere, so
it should not be confused with the 2D Toomre's parameter).
Using that at the midplane dw0/dz(0) = 0, we can derive
from equations (5-7) a single equation for the normalized
pseudo-enthalpy
2
s + 1
2γ (cid:16) dw0
dz (cid:17)
= (1 − w0) +
1
Q(s + 1)
(1 − ws+1
0
).
(9)
0, c2
The normalized density and sound speed are found from
ρ0 = ws
s = w0. Equation (9) shows that even though
the disc is in Keplerian rotation (i.e., self-gravity in the ra-
dial direction can be neglected), it must be included in de-
termining the vertical structure. We integrate equation (9),
starting at z = 0 with w0(0) = 1, until w0, monotonically
decreasing with z, reaches zero at some finite height/edge
z = h, which is therefore determined as a result of the inte-
gration process. At this edge, the density and sound speed
also vanish, ρ0(h) = 0, cs(h) = 0 (see also GLB and RPL).
The height and entire vertical structure of a polytropic disc
are therefore uniquely specified solely by Q and s, which are
free parameters in equation (9).
Figure 1 illustrates the sub- and superadiabatic equi-
librium vertical structures obtained from equation (9) for
various Q. For a fixed s, the disc height h decreases with
decreasing Q. The sound speed is a decreasing function of z
that does not permit the existence of two-dimensional modes
in three-dimensional models of polytropic discs as opposed
to isothermal ones (Lin et al. 1990, LP, KP). We also see
that the equilibrium structures for sub- and superadiabatic
cases look similar except that they have, respectively, every-
where positive and negative N 2
0 that diverges at the surface
because the sound speed vanishes there.
Figure 1. Vertical variations of the normalized density and
squared sound speed for subadiabatic (s = 5) and superadia-
batic (s = 1.7) equilibrium states with Q = 0.1 (thick solid lines),
Q = 0.5 (dashed-dotted lines), Q = 1 (dotted lines), Q = 10 (thin
solid lines) and the non-self-gravitating case (Q → ∞, dashed
lines). The lines corresponding to Q = 10 are close to those in the
non-self-gravitating limit, implying that the role of self-gravity is
negligible for Q > 10. The disc heights, h, are different depending
on Q and s.
2.2 Perturbation equations
We consider here small axisymmetric (∂/∂y = 0) perturba-
tions of the form u′, ρ′, p′, ψ ′ ∝ exp(−iωt + ikx) about the
equilibrium states found above, where ω and k are the fre-
quency and the radial wavenumber, respectively. Without
a loss of generality, we assume throughout that k is non-
negative, k > 0. After switching to non-dimensional vari-
ables:
t → Ωt, ω →
ω
Ω
,
k →
kcsm
Ω
, N0 →
N0
Ω
, g0 →
g0
Ωcsm
,
u′
x →
u′
x
csm
,
u′
y →
u′
y
csm
,
u′
z →
u′
z
csm
,
ρ′
ρm
p′
ρmc2
,
p′ →
, ψ ′ →
ρ′ →
we can derive from equations (1-4) the following set of equa-
tions governing the linear dynamics of axisymmetric pertur-
bations
sm
,
ψ ′
c2
sm
y − ikψ ′
ikp′
ρ0
− iωu′
− iωu′
− iωu′
+ 2u′
x = −
y = (q − 2)u′
dp′
dz − g0
z = −
1
ρ0
x
ρ′
ρ0 −
dψ ′
dz
− iωρ′ + ikρ0u′
x +
d
dz
(ρ0u′
z) = 0
− iω(p′ − c2
sρ′) +
ρ0c2
sN 2
g0
0
u′
z = 0
(10)
(11)
(12)
(13)
(14)
(cid:18)−k2 +
d2
dz2(cid:19) ψ ′ =
ρ′
Q
.
Axisymmetric modes in vertically stratified self-gravitating discs
5
(15)
p = g0uz
at
z = h.
(21)
If we now make the changes: iu′
z, ωp′/ρ0 → p′, ωψ ′ →
ψ ′ and eliminate u′
y, ρ′ from equations (10), (11), (13),
we arrive at the following set of equations for the three basic
quantities u′
z, p′, ψ ′ (henceforth primes will be omitted)
z → u′
x, u′
uz −(cid:18) 1
s −
c2
duz
dz
=
g0
c2
s
dp
dz
=
N 2
0
g0
p + (ω2 − N 2
Q (cid:18) p
c2
s
ρ0
d2ψ
dz2 − k2ψ =
k2
ω2 − κ2(cid:19) p +
0 )uz −
dψ
dz
+
N 2
0
g0
uz(cid:19) ,
k2ψ
ω2 − κ2
(16)
(17)
(18)
where the non-dimensional epicyclic frequency κ is given
by κ2 = 2(2 − q) = 1. Notice that the density pertur-
bations on the right hand side of the linearized Poisson's
equation (18) consist of two physically different parts. The
density perturbations due to compressibility, which are in-
versely proportional to the squared sound speed and pro-
portional to the pressure perturbation, and the density per-
turbations due to the stratified background, which are pro-
portional to the vertical velocity perturbation and N 2
0 . This
latter term due to stratification is obviously absent in the
two-dimensional analysis of gravitational instability as well
as when the disc is adiabatic (see GLB). Equations (16-
18) form the basis for determining the axisymmetric normal
modes of perturbations existing in compressible, stratified,
polytropic, self-gravitating discs. In the non-self-gravitating
limit, these equations reduce to the main equations of KP.
The reflection symmetry of the unperturbed disc allows us
to take the eigenfunctions/modes to be either even or odd
with respect to z. So, to determine the eigenfrequencies (dis-
persion diagrams, that is, ω as a function of k for various
branches) and eigenfunctions, we can numerically integrate
the main equations (16-18) only in the upper half, 0 6 z 6 h,
of the full vertical extent of the disc provided that suitable
boundary conditions are imposed at z = 0 and z = h. In
other words, we need to solve a boundary value problem.
Here the disc height h, as noted above, is determined by
the parameters Q and s. By inspecting equations (16-18),
it is clear that a normal mode whose pressure p and po-
tential ψ perturbations are even functions under reflection
in z, has odd vertical velocity uz and gravitational poten-
tial derivative dψ/dz perturbations under reflection and vice
versa. We define a normal mode as being 'even' ('odd') if
the corresponding vertical velocity and gravitational poten-
tial derivative are odd (even) functions of z (this convention
agrees with that of Ogilvie 1998, but differs from LP and
KP). Thus, the boundary conditions at the midplane, z = 0,
for the even modes are:
uz(0) = 0, dψ/dz(0) = 0
and for the odd modes are:
p(0) = 0, ψ(0) = 0.
(19)
(20)
At z = h we impose the usual free-surface boundary con-
dition for which the Lagrangian pressure perturbation van-
ishes. In our new non-dimensional variables this translates
as
The boundary condition for the gravitational potential can
be derived by treating the surface displacement as a surface
distribution of gravitating matter at z = h. Using the conti-
nuity of potential across the boundary, Gauss's flux theorem
and the fact that outside the disc, at z → ±∞, gravitational
potential should vanish, we arrive at the following condition
at the disc surface (see e.g., GLB, Lubow & Pringle 1993a)
dψ
dz
+ kψ = −
uzρ0
Q
at
z = h.
(22)
Equations (16-18) supplemented with boundary conditions
(21),(22) at the surface z = h and (19),(20) at the mid-
plane fully determine a boundary value problem. Using these
boundary conditions and variational principle, it can be eas-
ily shown that ω2 is real (GLB, RPL) that much alleviates
the search of eigenfrequencies. We integrate these equations
from z = h downwards to z = 0 separately for the even
modes under conditions (19) and for the odd modes un-
der conditions (20). We will see below that in the presence
of self-gravity, the dispersion diagrams for these two mode
parities are quite different. In other words, self-gravity in-
fluences even and odd modes, or changes their dispersion
characteristics, in different ways. As is typical, for a given
equilibrium structure, that is, for given Q and s, and for a
given radial wavenumber k, midplane boundary conditions
can be satisfied only for certain values of ω, yielding disper-
sion relations ω(k) for different branches of modes classified
below. We use a Runge-Kutta integrator and root-finding
algorithms of MATLAB package to first numerically solve
differential equations (16-18) and then find eigenfrequencies.
3 THE CLASSIFICATION OF VERTICAL
MODES
In this section, we for the moment remove self-gravity (put
ψ → 0, Q → ∞) from the perturbation equations (16-18)
and from the equilibrium in order to classify and character-
ize all the vertical axisymmetric normal modes existing in a
three-dimensional disc. In special cases, an analogous classi-
fication of modes in non-self-gravitating discs has been done
previously by several authors (RPL, LP, KP, Ogilvie 1998).
The aim here is to briefly review and synthesize the results
from these studies. In the next section, we will again switch
on self-gravity and investigate how it affects the characteris-
tics of these modes. Accordingly, in the non-self-gravitating
case, we need the midplane and surface boundary conditions
discussed above, but only for pressure and vertical velocity
perturbations.
Figure 2 shows the typical dispersion diagrams for three
types of equilibria: subadiabatic (1 + 1/s < γ), adiabatic
(1 + 1/s = γ) and superadiabatic (1 + 1/s > γ) stratifica-
tions. In the subadiabatic case, one can identify four distinct
classes of vertical normal modes (see also KP, Ogilvie 1998).
These modes are: the acoustic p-modes, the surface grav-
ity f-modes, the buoyancy g-modes and the inertial r-modes
the restoring forces for which at large radial wavenumbers
kh ≫ 1 are mainly provided, respectively, by compressibil-
ity/pressure, by the displacements of free surface of the disc,
by buoyancy due to the vertical entropy gradient and by
6 G. R. Mamatsashvili and W. K. M. Rice
(a)
50
40
ω2
30
20
10
1
0
8
(b)
ω2
6
4
2
1
0
2
4
2
4
6
k
8
10
12
(d)
60
40
ω2
20
1
0
60
(f)
40
ω2
20
1
0
2
4
2
4
6
k
6
k
8
10
12
8
10
12
6
k
ω2
1
0.8
0.6
0.4
0.2
0
0
1
ω2
0
−1
−2
0
(c)
1
0.8
0.6
0.4
0.2
ω2
8
10
12
0
0
2
4
8
10
12
6
k
(e)
2
4
6
k
8
10
12
(g)
2
4
6
k
8
10
12
5, po
4, po
4, pe
3, po
3, pe
2, po
2, pe
1, po
1, pe
Figure 2. Dispersion diagrams in the non-self-gravitating case for a subadiabatic disc with s = 5 (panels a, b, c), for an adiabatic
disc with s = 2.5 (panels d, e) and for a superadiabatic disc with s = 1.7 (panels f, g). The solid lines in panels (a),(d),(f) show the
frequencies for different branches of the surface gravity f- and acoustic p-modes vs. radial wavenumber k. In order of increasing frequency
these branches/curves are f e, f o, pe
5 (superscripts denote the even and odd modes, which are numbered,
respectively, according to the number of nodes of vertical velocity and pressure perturbations in the interval 0 < z 6 h). Although,
for large k, the frequencies of even and odd modes coalesce. The dashed lines in these panels correspond to frequencies of the f-modes
computed in the incompressible limit. The solid lines in panel (b) show the convectively stable g-mode branches, which in order of
decreasing frequency are go,e
5 . Here the frequencies of even and odd modes coincide. The solid lines in panels (c),(e)
show the r-mode branches, which in order of increasing frequency are ro
0 are the basic even and odd
r-modes and subscript '0' means that, respectively, vertical velocity and pressure perturbations for them have no nodes in the interval
0 < z 6 h. In this respect, our numbering of modes differs from that of Ogilvie 1998). The solid lines in panel (g) show the convectively
unstable g-modes, which in order of increasing ω2 are go
0 denotes the basic even g-mode, the corresponding
vertical velocity perturbation for which has no nodes in the range 0 < z 6 h). Similar to the p-modes, the convectively unstable even and
odd g-modes coalesce at large k. In each panel, the dashed lines show the corresponding mode branches computed for the incompressible
case with the same ordering of eigenfrequencies. From panels (b),(g), we see that compressibility most strongly affects the convectively
stable and unstable g-modes.
0 and ro
4 , go,e
3 (ge
1 , go,e
2 , go,e
3 , go,e
0, go
2, ge
2, go
4, ge
1, go
3, ge
1, ro
2, re
0, ro
1, re
1, ge
0, re
2, ro
3, re
3 (re
inertial forces due to disc rotation. The existence of the g-
modes in polytropic discs is attributable to the variation of
the sound speed and N0 with height; the latter diverges at
the disc surface giving rise to the trapped g-mode there. For
1 + 1/s < γ, all these modes have ω2 > 0 and, therefore,
are stable. The p-, f- and g-modes are high-frequency modes
with frequencies always larger than the epicyclic frequency,
ω2 > κ2, while the r-mode is of low-frequency with ω2 6 κ2
(we remind that hereafter κ = 1).
For adiabatic stratification, there is no buoyancy (N 2
0 =
0) and, therefore the g-mode disappears, while other modes
remain qualitatively unchanged. Similarly, the r-mode does
Axisymmetric modes in vertically stratified self-gravitating discs
7
not exist in a disc that has zero epicyclic frequency. For 1 +
1/s > γ, the g-mode becomes convectively unstable (ω2 < 0)
for radial wavenumbers larger than a certain value because
of negative buoyancy. Its characteristic timescale (growth
rate) can be of the order of the epicyclic frequency or less
and due to this, it interferes with motions corresponding to
the r-mode. Consequently, in the superadiabatic case, the r-
and g-modes merge at ω2 6 κ2 and appear as a single mode,
which we still call the g-mode and not the r-mode, because
the behaviour of corresponding eigenfunctions with height in
this case is similar to that of the convectively stable g-mode
(see also RPL); the p- and f-modes are not much affected.
All these modes come in even and odd pairs. In Fig. 2,
even (odd) modes are numbered according to the number
of nodes of the vertical velocity (pressure) perturbation in
the interval 0 < z 6 h (the node at z = 0 is not counted).
We also see the coalescence of the dispersion curves of the
even and odd p- and f-modes and also the convectively un-
stable even and odd g-modes as k increases past a certain
point, which depends on the mode number (RPL, KP). This
is associated with a transition from oscillatory to evanes-
cent behaviour of eigenfunctions. We do not plot eigenfunc-
tions here, as an extensive discussion of their properties
can be found in RPL, KP, LO98 and Ogilvie (1998). We
only mention that the eigenfunctions of the p- and g-modes
are trapped near the surfaces and decay towards the mid-
plane, while the eigenfunctions of the r-mode are concen-
trated near the midplane and decay towards the surfaces.
As for the eigenfunctions of the fundamental f-mode, they
have no nodes and monotonically decay from the surfaces
to the midplane. Below we show that these properties of
eigenfunctions are altered due to self-gravity. Specifically,
the number of nodes along each branch of the dispersion di-
agrams, which is preserved in the non-self-gravitating limit,
is no longer constant in the presence of self-gravity. The fre-
quencies of the p- and r-modes increase with mode number.
The frequencies and growth rates of the convectively sta-
ble and unstable g-modes, respectively, decrease with the
mode number. These are the well-known general properties
of vertical modes in polytropic discs.
Although each mode type has its dominant restoring
force, one out of the above mentioned four types (compress-
ibility, surface gravity, buoyancy, inertial forces), the other
three forces also contribute to a certain degree for small ra-
dial wavenumbers kh <
∼ 1. For example, the p- and f-modes
are modified both by rotation/inertial forces and buoyancy,
the f-mode is also modified by compressibility, the g-mode
is modified by rotation and compressibility and the r-mode
is modified by buoyancy and compressibility. For the sake
of comparison with the relevant result from previous stud-
ies, we are particularly interested in to what degree com-
pressibility modifies generally non-compressive f-, g- and r-
modes. So, we decided to juxtapose the dispersion diagrams
for these modes computed separately in the compressible
and incompressible cases. We take the incompressible limit
by formally letting the adiabatic index go to infinity, γ → ∞
(Ogilvie 1998). After that equations (16-18) without self-
gravity take the form
duz
dz
=
k2
ω2 − κ2 p
dp
dz
=
N 2
0
g0
p + (ω2 − N 2
0 )uz.
We solved these equations with the same boundary condi-
tions and found the dispersion diagrams of the f-, g- and
r-modes that survive in this limit (obviously, the acoustic
p-mode disappears). The results are plotted in Fig. 2 with
dashed lines. Notice that by taking the incompressible limit
in this manner, we have been able to retain the f-mode, as ex-
pected. Instead, KP set the density perturbations to zero in
the linearized continuity equation (anelastic approximation)
that resulted in the f-mode disappearing in their incompress-
ible limit. We see that compressibility most strongly affects
the convectively stable and unstable g-mode branches with
small mode numbers (Figs. 2b,2g), or equivalently with ver-
tical extent comparable to the disc height, even at large
radial wavenumbers (Ogilvie 1998). The frequencies of the
f- and r-modes do not change much, indicating that these
modes are nearly incompressible.
4 VERTICAL MODES IN THE PRESENCE OF
SELF-GRAVITY
Having characterized all types of axisymmetric normal
modes in a stratified disc, let us now compute the disper-
sion diagrams taking into account self-gravity in the per-
turbation equations and in the equilibrium. This will allow
us to understand how the frequencies and the structure of
the eigenfunctions of the above-described mode types are
altered by self-gravity, which mode acquires the largest pos-
itive growth rate in the presence of self-gravity and, thus
determines the onset criterion and nature of gravitational
instability of a disc. In other words, we return to the bound-
ary value problem formulated in Section 2, which is repre-
sented by equations (16-18) supplemented with boundary
conditions (21),(22) at the surface and conditions (19),(20)
of the even and odd symmetry of a solution at the midplane.
For each Q and s, we first determine the corresponding ver-
tical distribution of the equilibrium quantities in equations
(16-18), ρ0, c2
0 , with height as described in Section 2
and then based on this compute the normal modes.
s, g0, N 2
Figures 3,4,5 and 6 show the typical dispersion dia-
grams for the p-, f-, g- and r-modes in a self-gravitating
disc for subadiabatic, adiabatic and superadiabatic vertical
stratifications. We separately plot the dispersion diagrams
of the even and odd parities for each mode type to clearly
see the influence of self-gravity on them, which, as evident
from these figures, depends on the mode parity. Unlike in
the non-self-gravitating limit, the number of nodes of the
vertical velocity and pressure perturbations in the presence
of self-gravity are not preserved along each mode branch.
However, as we will see below, at large k, the influence of
self-gravity on the mode dynamics is small and the disper-
sion diagrams merge with their non-self-gravitating coun-
terparts (shown with dashed curves in Figs. 3, 5). So, the
naming of the modes for smaller k, where the effect of self-
gravity is important, is done by continuity with the large-k
limit for each mode branch.
8 G. R. Mamatsashvili and W. K. M. Rice
80
60
ω2
40
20
1
0
2
80
60
ω2
40
20
1
0
2
(b)
(e)
ω2
ω2
2.6
2.4
2.2
2
1.8
1.6
1.4
1.2
1
0
2.6
2.4
2.2
2
1.8
1.6
1.4
1.2
1
0
(a)
6
8
(d)
6
8
4
k
4
k
ω2
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
(c)
2
k
4
0
2
4
6
k
ω2
1
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
(f)
2
k
4
0
2
4
6
k
Figure 3. Dispersion diagrams of vertical modes in a subadiabatic self-gravitating disc with s = 5 and Q = 0.2. Shown are (a) the even
p- and f-modes, (b) the even g-mode, (c) the even r-mode, (d) the odd p- and f-modes, (e) the odd g-mode, (f) the odd r-mode. Dashed
lines are the corresponding branches computed without self-gravity in the perturbations, but with the same self-gravitating equilibrium.
The frequency ordering and mode naming are the same as in Fig. 2. Self-gravity reduces/shifts the frequencies of the even and odd p-,
f-modes and the even g- and r-modes mainly in the range 0 6 k 6 2; the frequencies of the odd g- and r-modes are almost unaffected by
self-gravity. The frequency of the basic even r-mode (denoted above as re
0) is modified most strongly by self-gravity as evidenced by the
largest dip on the corresponding dispersion curve in panel (c). For large k and/or large mode numbers, the effect of self-gravity becomes
weak and the dispersion curves merge with the dashed ones for the non-self-gravitating case.
4.1 Subadiabatic stratification
Consider first the subadiabatic case (Fig. 3), where we again
have the p-, f-, g- and r-modes, but their dispersion diagrams
are modified/shifted from their non-self-gravitating counter-
parts towards lower values due to self-gravity. As illustrated
in Figs. 3a,3d, self-gravity reduces the frequencies of all
branches of the p- and f-modes, for both even and odd pari-
ties, but it more affects the f-modes. The situation for the g-
and r-modes is different (Figs. 3b,3c,3e,3f): only the frequen-
cies of the even g- and r-modes are reduced by self-gravity
mostly for radial wavenumbers in the range 0 6 k 6 2 (dips
on the corresponding dispersion curves in Figs. 3b,3c indi-
cating deviations from the non-self-gravitating dashed ones),
whereas the frequencies of the odd g- and r-modes are al-
most unaffected by self-gravity. The frequencies of the fun-
damental f-modes, the first few branches of the p-modes and
also the first few branches of the even g- and r-modes are
reduced noticeably. With increasing mode number and/or
radial wavenumber, the effect of self-gravity on the eigenfre-
quencies gradually falls off, because the corresponding eigen-
functions become of smaller and smaller horizontal and/or
vertical scale (the number of nodes increases) and take the
form similar to that in the non-self-gravitating limit. As
a result, the dispersion diagrams become more and more
identical to their non-self-gravitating counterparts. The or-
dering of frequencies for the even and odd p-, f-, g- and
r-modes, as it is in the non-self-gravitating case (see Fig. 2):
ω2(r) 6 κ2 6 ω2(g) < ω2(f) < ω2(p), is preserved in the
self-gravitating case as well, however small the value of Q.
Note also in Fig. 4 that the frequencies of the the even and
odd f-modes and also of the basic even g-mode, for which
the influence of self-gravity is stronger, never fall below the
epicyclic frequency with decreasing Q and, therefore, these
modes always remain gravitationally stable (we also checked
that the same situation holds for the dispersion curves of the
even and odd f-modes in the adiabatic and superadiabatic
cases described below). This is a consequence of the three-
dimensionality of the disc. Thus, as evident from Figs. 3c,4d,
self-gravity most significantly affects the basic branch of the
Axisymmetric modes in vertically stratified self-gravitating discs
9
Q=0.2
Q=0.05
Q=0.01
60
40
20
ω2
60
40
20
ω2
1
0
2
4
(a)
8
10
12
6
k
1
0
2
4
(b)
8
10
12
6
k
ω2
12
10
8
6
4
2
1
0
ω2
0
−5
−10
−15
(c)
2
4
6
k
8
10
12
−20
0
2
4
(d)
8
10
12
6
k
Figure 4. Dispersion curves for s = 5 and Q = 0.01 (dashed-
dotted lines), Q = 0.05 (dashed lines), Q = 0.2 (solid lines). Pan-
els (a),(b) show the even and odd f-modes, respectively. Panel (c)
shows the basic even g-mode and panel (d) shows the basic even
r-mode. The frequencies of the f- and g-modes never fall below
the epicyclic frequency with decreasing Q and, therefore, are al-
ways stable. The basic even r-mode, which is most significantly
affected by self-gravity, becomes unstable when the dip on the
corresponding dispersion curve crosses the ω2 = 0−axis. This dip
deepens and broadens with decreasing Q.
even r-mode and only this branch can become unstable due
to self-gravity. From Fig. 4d, it is seen that the dip on this
branch deepens and broadens with decreasing Q. (We de-
note the wavenumber at which the minimum of this dip is
located by km.) The gravitational instability sets in when
the dip's minimum first touches the ω2 = 0−axis at some
Qcr and then farther extends into the unstable ω2 < 0 region
for Q < Qcr. We will demonstrate in Section 5 that the basic
even r-mode becomes strongly compressible at k ∼ km (Fig.
8), as opposed to the r-mode in a non-self-gravitating disc,
and the density perturbations due to compressibility are re-
sponsible for its gravitational instability. For s = 5, we find
Qcr = 0.168, which gives H = 2h = 2.96 for the disc total
thickness, and km = 0.8 (see Fig. 12). Note that the charac-
teristic radial scale of instability λm = 2π/km = 7.85 is not
much greater than the disc thickness H. In self-gravitating
discs, compressibility and inertial forces play an important
role in the dynamics of the basic even r-mode at wavelengths
λ ∼ λm
∼ H and so it resembles a 2D density wave at such
wavelengths (see Section 4.4). The fact that the character-
istic scale of gravitational instability λm turns out to com-
parable to the disc thickness may offer a clue why angular
momentum transport due to self-gravity tends to be a lo-
cal phenomenon (e.g., Gammie 2001; Lodato & Rice 2004;
Boley et al. 2006). In this respect, it should also be men-
tioned that the analogous gravitational instability of low-
frequency modes in a strongly compressed gaseous slab was
also found by Lubow & Pringle (1993a). In their analysis,
these modes, called neutral modes, are basically degenerate
r-modes, since the compressed gaseous slab was not rotating.
>
4.2 Adiabatic stratification
In the adiabatic case (Figs. 5a-5d), there is no g-mode and
the behaviour of the p-, f- and r-modes under self-gravity
has qualitatively the same character as in the subadiabatic
case above. Again, the basic branch of the even r-mode ap-
pears to be most significantly affected by self-gravity (the
corresponding dispersion curve in Fig. 5c has the same form
as that in Fig. 3c) and becomes gravitationally unstable.
The ω2 of the p- and f-modes, although lowered by self-
gravity, always remain larger or equal to κ2 irrespective of
the value of Q. In this case, s = 2.5 and the corresponding
Qcr = 0.177 and km = 0.81. Again the radial scale of insta-
bility λm = 2π/km = 7.75 is not far from the disc thickness
for these parameters, H = 2h = 2.08.
4.3 Superadiabatic stratification
The superadiabatic case (Figs. 5e-5h) is interesting, because
at ω2 6 κ2, as classified above, instead of the r-mode there
is the g-mode, which in addition to convective instability
can also exhibit gravitational instability. As in the non-self-
gravitating case, for radial wavenumbers larger than a cer-
tain value, all the branches of the g-mode become unstable
because of the negative vertical entropy gradient. From Fig.
5g we see that for 0 6 k 6 2, self-gravity produces dips on
the g-mode dispersion curves, again preferably for the even
parity ones. When Q drops sufficiently, the dip on the ba-
sic branch of the even g-mode, which appears to be most
affected by self-gravity, starts to cross the ω2 = 0-axis in a
way similar to that of the basic even r-mode branch above
and this signals the onset of gravitational instability. For
s = 1.7, adopted in Fig. 5, we find Qcr = 0.184, which
gives a thickness H = 2h = 1.76, and km = 0.83, so that
λm = 2π/km = 7.57 >
∼ H. As for the p- and f-modes, they
behave under self-gravity in exactly the same manner as in
the above cases. In particular, although their frequencies
are reduced, they can never become gravitationally unsta-
ble even for very small values of Q. Thus, for superadiabatic
stratification, the basic even g-mode can exhibit two types
of instabilities: gravitational and convective (see Fig. 6). As
we see from Figs. 5g,6a, at moderate Q ∼ Qcr, the radial
scales for the activity of self-gravity and convective insta-
bility are well separated, so that these two effects do not
interfere with each other in the linear regime. For Q = 0.1
in Fig. 6a, self-gravity is dominant at 0 6 k 6 2.55, while
convective instability occurs at k > 4.27, where the effect of
self-gravity is weak (a similar situation is for Q = 0.2 in Fig.
5g). However, for very small Q ≪ Qcr, the radial scales of
gravitational and convective instabilities overlap (Fig. 6b),
but the gas motion for such radial scales in the case of very
strong gravitational instability would hardly resemble con-
vective motions (it will look more like that in Fig. 11 below);
convection is simply disrupted by gravitational instability.
Therefore, we can conclude that unless a disc is strongly self-
gravitating, self-gravity has little influence on the properties
of convective motions and on the (Schwarzschild) criterion
for the onset of convective instability that is equivalent to
N 2
0 < 0.
10 G. R. Mamatsashvili and W. K. M. Rice
ω2
ω2
80
60
40
20
1
0
80
60
40
20
1
0
80
60
ω2
40
20
(a)
10
1
0
80
60
ω2
40
20
(e)
10
1
0
5
k
5
k
5
k
5
k
1
0.8
0.6
ω2
0.4
0.2
(b)
10
0
0
(c)
4
2
k
1
0.5
ω2
0
−0.5
1
0.8
0.6
ω2
0.4
0.2
0
0
1
0.5
ω2
0
−0.5
(d)
4
2
k
(f)
−1
(g)
−1
10
0
2
4
k
6
8
0
2
(h)
6
8
4
k
Figure 5. Dispersion diagrams for vertical modes in a self-gravitating disc with Q = 0.2 for an adiabatic stratification with s = 2.5
(panels a-d) and for a superadiabatic stratification with s = 1.7 (panels e-h). Shown are (a,e) the even p- and f-modes, (b,f) the odd p-
and f-modes, (c) the even r-mode, (d) the odd r-mode, (g) the convectively unstable even g-mode and (h) the odd g-mode. The frequency
ordering and mode naming are the same as in Fig. 2. As before, dashed curves correspond to the dispersion diagrams calculated without
self-gravity in the perturbations. The behaviour of the p-, f- and r-modes in the adiabatic case is similar to that in Fig. 3. In the
superadiabatic case, the dispersion curves of the even g-mode have dips caused by self-gravity in the interval 0 6 k 6 2, while the
odd g-mode is almost unaffected by self-gravity. The largest dip occurs for the basic branch of the even g-mode (denoted above as ge
0)
in panel (g), which appears to be most significantly influenced by self-gravity. For large k and/or large mode numbers, the effect of
self-gravity becomes weak and the dispersion curves merge with dashed ones for the non-self-gravitating case. Due to the superadiabatic
stratification, the even and odd g-modes are convectively unstable (i.e., have ω2 < 0) in the range k > 3.91, where the influence of
self-gravity is small.
4.4 Analogy with the 2D density wave theory
and the 2D density wave theory in thin self-gravitating discs,
which actually appears to do a decent qualitative job.
From the three cases considered above, we conclude that
the basic branch of the low-frequency (ω2 6 κ2) even mode,
which is the r-mode, in the case of subadiabatic and adia-
batic stratifications, and the g-mode, in the case of superadi-
abatic stratification, is most subject to the influence of self-
gravity. These modes determine the gravitational instability
in vertically stratified discs. As we have clearly seen in all the
above cases, the radial scale of the instability is comparable,
though a bit larger than the disc thickness, λm
∼ H, which
in turn implies that a 2D analysis of the gravitational insta-
bility is at most marginally valid and the rigorous stability
(linear) analysis of self-gravitating discs should be three-
dimensional. However, despite the 2D treatment only being
marginal, we can still find similarities between our results
>
Consider an equivalent zero-thickness disc with the
ρ0dz.
sound speed csm
Then the well-known dispersion relation of axisymmetric 2D
density waves in the thin disc is (e.g., Goldreich & Tremaine
1978)
1 and the surface density Σ0 = R h
−h
2
Q2D
ω2 = k2 −
where we have used the same normalization as before and
k + 1,
1 In fact, there are other options for choosing the sound speed as
a some kind of height-averaged value. We address this question in
Section 6, but for the present purpose this does not make much
difference.
Axisymmetric modes in vertically stratified self-gravitating discs
11
(a)
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
−1
ω2
0
2
4
k
6
8
ω2
1
0
−2
−4
−6
−8
−10
−12
−14
−16
−18
−20
0
(b)
5
k
10
15
k
m
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0.1
0.2
0.3
0.4
0.5
0.6
Q
Figure 6. Dispersion curves of the basic even g-mode for a su-
peradiabatic stratification with s = 1.7 for (a) Q = 0.1 and
(b) Q = 0.01. Dashed lines show the same branches computed
without self-gravity in the perturbations. For Q = 0.1, the basic
even g-mode exhibits gravitational instability in the vicinity of
km = 1.15 and convective instability in the range k > 4.27, where
the influence of self-gravity is weak. For Q = 0.01, the effect of
self-gravity is much stronger and the radial scales of gravitational
and convective instabilities overlap in the range 4.25 6 k 6 8.5.
because of that Q2D = csmΩ/πGΣ0. This dispersion rela-
tion is a parabola with a minimum at the Jeans wavenum-
ber kJ = 1/Q2D. This is the wavenumber, at which the
effect of self-gravity is most prominent, and if Q2D < 1
it gives the characteristic scale of gravitational instability.
At small k ≪ kJ , 2D density waves are dominated by self-
gravity and inertial forces, while at large k ≫ kJ , pres-
sure/compressibility dominates over self-gravity and density
waves appear as an acoustic mode. At k ∼ kJ all three fac-
tors can be important. Let us now look at the dispersion
curves of the basic branches of the even r- and convectively
unstable even g-modes in Figs. 3c,5c,5g. They have similar
parabolic shape in the self-gravity and compressibility dom-
inated regime at k ∼ km with the linear phase at smaller
k ≪ km, where only self-gravity and inertial forces play
a role. This linear phase at long wavelengths is well repro-
duced by the 2D dispersion relation. Therefore, we can iden-
tify the wavenumber km, at which the effect of self-gravity
on the 3D modes is largest, with the Jeans wavenumber kJ .
In some sense, in the case of instability when Q < Qcr, km
gives a more accurate value for the radial scale, λm, of the
gravitationally most unstable mode than that given by the
Jeans wavenumber kJ in the 2D model. For example, from
Fig. 12 we find that the vertical structure (ρ0) calculated at
s = 5 and Qcr = 0.168 gives the corresponding Q2D = 0.76
and kJ = 1.32, which differs from km = 0.8 found above for
these parameters. Moreover, we will see in Section 6 that the
criterion for gravitational instability based on the 3D calcu-
lations, i.e. Q < Qcr, is more exact than the 2D criterion
Q2D < 1.
In the 2D case, as shown above, the Jeans wavenumber
kJ = 1/Q2D, implying that it is determined solely by com-
petition between self-gravity and compressibility/pressure.
It is now interesting to see how an analogous characteristic
Figure 7. Dependence of km on Q for s = 5, which very closely
follows the power law Q−1/2 (shown with dashed line and scaled
appropriately). The gravitational instability sets in at Q = Qcr =
0.168 that gives km = 0.8.
wavenumber km depends on Q in the 3D case. From Fig.
7 we see that this dependence has a power law character
Q−1/2, which means that the value of km is again set by
self-gravity and compressibility, as in the 2D case, so that
the disc rotation plays a role only in determining ω2 but not
km. Indeed, the only lengthscale that can be constructed
from the sound speed csm, density ρm and gravitational con-
stant G without the angular velocity, Ω, of disc rotation is
csm/√Gρm (analogous to the Jeans length of a collapsing
3D cloud). If we normalize it by csm/Ω, we get 2√πQ1/2,
giving the above power law dependence for the correspond-
ing non-dimensional wavenumber √πQ−1/2 to which km is
proportional.
Thus, the basic even r-mode or, in the case of superadia-
batic stratification, the basic even g-mode at k <
∼ km exhibit
many of the characteristics of 2D density waves and could be
regarded as their 3D analogues at such radial wavenumbers
(see also Section 5). If continued to k ≫ km, the density
wave mode would thus connect up to the large-k parts of
these two modes dominated, respectively, by inertial forces
and by negative buoyancy, instead of merging with the com-
pressible p-mode, as might seem at first sight from the 2D
dispersion relation. However, the limit k ≫ km, equivalent
to λ ≪ H, means the breakdown of the 2D approximation
and, therefore, the concept of 2D density waves at k ≫ km
would not be well-defined.
5 GRAVITATIONAL INSTABILITY IN 3D:
PROPERTIES OF THE BASIC BRANCH OF
THE LOW-FREQUENCY EVEN R-MODE
In this section we concentrate only on the properties of the
gravitationally unstable basic even r-mode. As we have seen,
the behaviour of the convectively unstable basic even g-mode
under self-gravity has a qualitatively similar character.
12 G. R. Mamatsashvili and W. K. M. Rice
ω2
1
0.8
0.6
0.4
0.2
0
−0.2
−0.4
−0.6
−0.8
0
1
2
3
k
4
5
6
Figure 8. Dispersion curves of the basic even r-mode at s =
5, Q = 0.1 with corresponding h = 1.22. Dotted line corre-
sponds to the non-self-gravitating case, dashed-dotted -- to the
incompressible case with self-gravity and solid line -- to the self-
gravitating case with compressibility. The influence of self-gravity
on the basic even r-mode is very weak in the incompressible limit
and the dashed-dotted dispersion curve almost coincides with the
dotted one in the non-self-gravitating limit; a slight deviation
(dip) at small k is due to the stratified background.
5.1 Effect of compressibility
As noted above, in the Poisson's equation, the density per-
turbations giving rise to the gravitational potential pertur-
bations can be due to compressibility and due to the strat-
ification of the equilibrium vertical structure. As is evident
from Fig. 2c, in non-self-gravitating discs, the r-mode is
nearly incompressible. So, it is interesting to see if it still re-
mains incompressible in self-gravitating discs and which out
of these two sources of density perturbations is ultimately
responsible for the gravitational instability. In order to ex-
plore this, we again computed the dispersion curve of the
basic even r-mode in the incompressible limit of equations
(16-18), which take the form:
k2
duz
dz
=
dp
dz
=
ω2 − κ2 (p + ψ)
p + (ω2 − N 2
N 2
0
g0
dψ
dz
(23)
(24)
0 )uz −
uz.
(25)
ρ0
Q
N 2
0
g0
d2ψ
dz2 − k2ψ =
supplemented with the same boundary conditions. Notice
that on the right hand side of Poisson's equation (25),
only the density perturbation due to stratification remains.
In Fig. 8, we compare the dispersion curves of the basic
even r-mode obtained in the incompressible limit to those
computed for the compressible self-gravitating and non-self-
gravitating cases. It is clear that in self-gravitating discs, the
basic even r-mode becomes compressible at k ∼ km thereby
providing density perturbation for the gravitational poten-
tial and, therefore, accounting for the large dip on the dis-
persion curve. The dispersion curve in the incompressible
limit deviates only very slightly from that in the non-self-
gravitating case, because of the second source of density
20
0
u
x
−20
−40
−60
0
20
0
p
−20
−40
−60
0
u
z
ρ
1.5
1
0.5
0
0
20
0
−20
−40
−60
0
0.5
z
0.5
z
1
1
0.5
z
0.5
z
1
1
Figure 9. Vertical structures of the radial velocity ux, the ver-
tical velocity uz, the pressure p and the density ρ perturbations
constituting the eigenfunctions of the gravitationally unstable ba-
sic even r-mode for s = 5, Q = 0.1, h = 1.22, km = 1 with the
corresponding largest growth rate ω2
min = −0.75. Dashed lines
show the eigenfunctions (scaled arbitrarily for plotting purposes)
in the non-self-gravitating case for the same parameters except
ω2 = 0.99. Notice that in the non-self-gravitating case, the eigen-
functions are more concentrated near the midplane, whereas in
the self-gravitating case they vary over the whole vertical extent.
perturbation, i.e. stratification, which turns out to be much
smaller than that associated with compressibility. Such a
comparison also shows that the primary cause of gravita-
tional instability in incompressible discs, as described in
GLB and Lubow & Pringle (1993a), is the displacements of
free-surfaces. In our case, the equilibrium density vanishes
at the surface, so that the effect of surface displacements on
the growth rate of instability is null (see boundary condition
22) and in the incompressible case only density perturbation
due to stratification, as shown, is hardly sufficient for grav-
itational instability.
5.2 Eigenfunctions and spatial structure
Here we compute the vertical structure of the eigenfunctions
of the gravitationally unstable basic even r-mode and also
find what type of motions it induces. We take Q = 0.1 and
s = 5, which gives h = 1.22 for the disc height. For these
parameters, the dispersion curve of this mode in Fig. 8 has
a minimum ω2
min = −0.75, i.e. gives the largest growth rate
of gravitational instability, at km = 1. In Fig. 9, we plot the
corresponding eigenfunctions only in the upper half of disc's
full vertical extent. Because the mode has even parity, the
vertical velocity and the derivative of the gravitational po-
tential are odd functions and, correspondingly, the pressure
and potential are even functions of z. Only the pressure per-
turbation has one node in the interval 0 < z 6 h, other quan-
tities have no nodes and vary over the whole vertical extent.
This demonstrates that the spatial structure of the gravita-
tionally most unstable mode is three-dimensional with non-
zero vertical velocity. To see how self-gravity changes the
form of the eigenfunctions, we also show the eigenfunctions
of the same branch in the non-self-gravitating limit, which
Axisymmetric modes in vertically stratified self-gravitating discs
13
u
x
p
60
40
20
0
−20
0
5
4
3
2
1
0
0
1
0.5
0
−0.5
0
0.04
0.02
0
u
z
ρ
−0.02
0
0.5
z
0.5
z
1
1
0.5
z
0.5
z
1
1
Figure 10. Eigenfunctions of the even f-mode, which unlike the
basic even r-mode is gravitationally stable, with the same values
of parameters as in Fig. 9. The corresponding eigenfrequencies at
km = 1 are ω2 = 1.1 for the self-gravitating case (solid lines) and
ω2 = 1.6 for the non-self-gravitating case (dashed lines). Again,
non-self-gravitating eigenfunctions have been scaled arbitrarily
for plotting purposes.
appear to be more concentrated towards the midplane than
the self-gravitating ones and have no nodes in the same ver-
tical range. Since the f-mode plays an important role in the
angular momentum and energy transport in stratified discs
(LO98), in Fig. 10 we also plot its eigenfunctions in the
presence of self-gravity for the same parameters. Compar-
ing Figs. 9 and 10, we clearly see that self-gravity modifies
the form of the eigenfunctions of both the f- and r-modes.
The perturbed quantities for both mode types vary with
height somewhat similarly, especially the vertical velocities,
which are, however, still at least an order of magnitude less
than the horizontal ones, so that the motions can be viewed,
to leading order, as 2D (see also Fig. 11). Thus, the gas mo-
tion associated with the gravitationally unstable r-mode is
of similar type to that of the f-mode, which, in turn, implies
that the r-mode might be as important as the f-mode, i.e.,
might be able to do similar 'dynamical jobs' as the f-mode
in self-gravitating discs. As mentioned earlier, the non-linear
behaviour of the 3D perturbations in self-gravitating discs
has been attributed solely to the f-mode dynamics without
analysing other modes (Pickett et al. 2000; Boley & Durisen
2006).
The
have
the
quantities
perturbed
form
f (z)exp(ikmx), f ≡ (ux, uz, p, ρ) and from this we can
construct the spatial picture of the velocity, density and
temperature perturbations for the gravitationally unstable
basic even r-mode by taking the real parts. The temperature
perturbation is found from the ideal gas equation of state
and is equal to T = −c2
sρ/ρ0 + γp/ρ0 (here the pressure
perturbation p is as used in the original equations 10-15,
i.e., before switching to new variables). Figure 11 shows
the density and temperature fields constructed in this way
with velocity vectors showing the direction of gas motion
superimposed on the density field, which in fact resembles
the classical density profile of a 2D density wave. Near the
midplane, matter flows (converges), almost parallel to the
Figure 11. Distribution of the density (upper plot) and temper-
ature (lower plot) perturbations in (x, z) plane constructed from
the eigenfunctions of the gravitationally unstable basic even r-
mode in Fig. 9. Superimposed on the density field are the velocity
vectors of the induced gas flow.
x−axis, towards high density regions. With increasing z,
the flow becomes more arc-like, because of the variation of
the vertical velocity with height. An analogous velocity field
in the (x, z) plane was also observed in some non-linear
simulations of self-gravitating discs (Boley & Durisen 2006;
Boley et al. 2006), which suggests that the latter may be a
result of the non-linear development of the type of motion
we see here in the linear regime. We might also speculate
that in relatively high-mass discs, the non-axisymmetric
gravitationally unstable basic r-mode will produce similar
high-density regions, which can be precursors of spiral
shock fronts (if the disc does not fragment), and arc-like
motions in the upper layers that can give rise to non-linear
shock bores involving large distortions of the disc surface
as described by Boley & Durisen (2006). The temperature
perturbations are fairly non-uniform as well, varying both
vertically and radially on comparable scales. Obviously,
larger temperature perturbations correspond to overdense
regions that are contracting because matter flows into
them, and lower -- to underdense regions.
6 STABILITY CRITERIA IN 2D AND 3D
In the 3D case, the degree and effect of self-gravity is charac-
terized by Q3D = Ω2/4πGρm, where ρm is the value of equi-
librium density at the midplane (see also GLB), which plays
here a role similar to that of the standard 2D Toomre's pa-
rameter Q2D = csΩ/πGΣ (Toomre 1964). However, in three-
dimensional simulations, researchers often still tend to use
the 2D Toomre's parameter to characterize the onset of grav-
itational instability (Rice et al. 2003, 2005; Lodato & Rice
2004, 2005; Boley et al. 2006). In this section we investigate
how a more general and exact 3D criterion of gravitational
instability can be related to the 2D one. In general, these
14 G. R. Mamatsashvili and W. K. M. Rice
1
0.8
0.6
0.4
0.2
0
2
1.5
1
0.5
Q
1 cr, 2D
Q
2 cr, 2D
Q
cr, 3D
1
2
3
4
k
J 2
k
J 1
k
m
1
2
3
4
5
s
5
s
6
7
8
9
10
6
7
8
9
10
Figure 12. Dependence of the critical Qcr,3D and the corre-
sponding critical Q1 cr,2D and Q2 cr,2D on the polytropic index
s. Shown also are the km at Qcr,3D and the 2D Jeans wavenum-
bers kJ 1 = 1/Q1 cr,2D, kJ 2 = 1/Q2 cr,2D as a function of s.
two parameters are different, because the 2D parameter con-
tains the sound speed and surface density, which by defini-
tion do not vary vertically for razor-thin discs, whereas the
3D parameter does not contain the sound speed and sur-
face density, but the midplane value of the density. So, if
we still want to characterize the 3D instability in terms of
the 2D Toomre's parameter, we need to use some height-
averaged, or effective sound speed (surface density can be
uniquely calculated from the vertical density distribution).
In some simulations the midplane values (Mej´ıa et al. 2005;
Boley et al. 2006), while in others vertically averaged values
(Rice et al. 2003, 2005) of the sound speed are used. Here we
calculate the corresponding critical Qcr,2D for these cases.
As is well-known, in razor-thin 2D discs, axisymmetric
perturbations (density waves) are gravitationally unstable if
Q2D < 1 (Toomre 1964). Let us now find the critical Qcr,3D
that determines the gravitational instability in 3D, which,
as found in this study, is associated with the basic branches
of the even parity low-frequency modes. In Fig. 12, we show
the critical Qcr,3D as a function of the polytropic index s.
Given Qcr,3D, we can now find Q1 cr,2D defined in terms of
the midplane sound speed and Q2 cr,2D defined in terms of
the vertically averaged sound speed:
Q1 cr,2D =
, Q2 cr,2D =
2Qcr,3D
0 ρ0dz
R h
0 csdz
,
2Qcr,3DR h
hR h
0 ρ0dz
where ρ0 and cs, as before, are the normalized equilibrium
density and sound speed the z-dependence of which, as well
as the value of h, are determined by Qcr,3D and the poly-
tropic index s. So, the critical Q1 cr,2D and Q2 cr,2D calcu-
lated with the two different methods are different, though
not far from each other, and less than unity which is their
critical value in the 2D case. This again confirms the well-
known result that 3D discs are more stable and to make
them unstable smaller value of Q2D is necessary when com-
pared with that required for razor-thin 2D discs (see also
GLB). But in such cases, using Q3D seems more appropri-
ate as it does not involve uncertainties in how to average the
sound speed over height.
In Fig. 12 we also examine the dependence of the crit-
ical wavenumbers of the gravitational instability on s. We
see that the 2D Jeans wavenumbers are always larger than
the actual more exact wavenumber km at the onset of grav-
itational instability when Q = Qcr,3D. Thus, in the 3D
case, the critical wavenumbers depend differently on the
thermodynamics than in the 2D case, where the critical
kJ = 1/Q2D = 1 at the onset of instability.
7 SUMMARY AND DISCUSSIONS
In this paper, we have analysed the axisymmetric nor-
mal modes of perturbations in stratified, compressible, self-
gravitating gaseous discs with subadiabatic, adiabatic and
superadiabatic vertical stratifications. First, we performed
a classification of perturbation modes in stratified discs in
the absence of self-gravity to compare with previous calcu-
lations. Four well-known main types of modes can be distin-
guished: acoustic p-modes, surface gravity f-modes, buoy-
ancy g-modes and inertial r-modes. The restoring forces
for these modes for large radial wavenumbers are mainly
provided by one the following: pressure/compressibility, dis-
placements of the disc surface, buoyancy and inertial forces
due to disc rotation for the p-,f-,g- and r-modes, respec-
tively. For smaller wavenumbers, the restoring force for each
mode is a combination of these forces. In the case of adi-
abatic stratification, buoyancy is zero and, therefore, the
g-mode disappears, while other modes remain qualitatively
unchanged. For superadiabatic stratification, the g-mode be-
comes convectively unstable and merges with the r-mode,
so that only a single convectively unstable mode appears in
the dispersion diagram at ω2 6 κ2, which we still call the
g-mode. Due to the reflection symmetry of the equilibrium
vertical structure with respect to the midplane, each mode
comes in even and odd pairs. By our terminology, for even
(odd) modes, pressure and gravitational potential pertur-
bations are even (odd), while the perturbations of vertical
velocity and derivative of potential are odd (even) functions
of the vertical coordinate. After classifying and character-
izing modes in the absence of self-gravity, we introduced
self-gravity in the perturbation equations and investigated
how it modifies the properties of these modes. We found
that self-gravity, to some extent, reduces the frequencies of
all normal modes at radial wavelengths comparable to the
disc height, but its influence on the basic even r-mode, in
the case of subadiabatic and adiabatic stratifications, and on
the basic even g-mode, in the case of superadiabatic stratifi-
cation, appears to be strongest. With decreasing Q3D, these
modes become unstable due to self-gravity and thus deter-
mine the gravitational instability of a vertically stratified
disc. The basic even g-mode also exhibits convective insta-
bility due to a negative entropy gradient but, unless the disc
is strongly self-gravitating, these two instabilities grow con-
currently in the linear regime, because their corresponding
radial scales are separated. We also obtained the correspond-
ing criterion for the onset of gravitational instability in 3D,
which is more exact than the standard instability criterion in
terms of the 2D Toomre's parameter, Q2D < 1, for axisym-
metric density waves in razor-thin discs. By contrast, the
Axisymmetric modes in vertically stratified self-gravitating discs
15
p-, f- and convectively stable g-modes have their ω2 reduced
by self-gravity, but never become unstable for any value of
Q3D. This is a consequence of the three-dimensionality of the
disc. The eigenfunctions associated with the gravitationally
unstable modes are intrinsically three-dimensional, that is,
have non-zero vertical velocity and all perturbed quantities
vary over the whole vertical extent of the disc. In this regard,
we would like to mention that resolving the gravitationally
most unstable mode in numerical simulations thus reduces
to properly resolving the disc height (together with resolv-
ing the corresponding radial scale, which, as shown here,
appears somewhat larger than the disc height). So, the cri-
terion of Nelson (2006) that at least ∼ 4 particle smoothing
lengths should fit into per scale height may apply in three-
dimensional SPH simulations. This implies a substantially
larger number of SPH particles per vertical column because
the disc itself may extend over many scale heights. He also
shows that a similar criterion applies to grid-based simula-
tions.
Here for simplicity, and also to gain the first insight into
the effects of self-gravity on the vertical modes in stratified
discs, we considered only axisymmetric perturbations. Non-
axisymmetric perturbations are dynamically richer, though
more complicated, because the phenomena induced by Ke-
plerian shear/differential rotation -- strong transient am-
plification of perturbations and linear coupling of modes
(not to be confused with non-linear mode-mode interac-
tions) -- come into play for these type of perturbations. Tran-
sient (swing) amplification of perturbations (density waves)
has been studied previously in razor-thin 2D approximation
(Goldreich & Lynden-Bell
1965b; Goldreich & Tremaine
1978; Toomre 1981; Mamatsashvili & Chagelishvili 2007).
From the analysis presented here, we may expect that
in the linear regime, the non-axisymmetric basic even r-
mode can undergo larger transient amplification due to self-
gravity than other modes in the disc. This transient am-
plification of perturbations may be important for explain-
ing the large burst phases seen in numerical simulations
at the initial stages of the development of gravitational
instability in discs (Rice et al. 2003, 2005; Lodato & Rice
2004, 2005; Boley et al. 2006). So, in this respect one should
analyse and quantify the transient amplification of non-
axisymmetric perturbations in stratified 3D self-gravitating
discs starting with linear theory. As for the linear coupling
of modes, it was demonstrated that in non-self-gravitating
stratified discs, Keplerian shear causes rotational (vortex)
mode perturbations to couple with and generate g-mode
perturbations (Tevzadze et al. 2008). In the context of 2D
discs, it was shown that vortex mode perturbations can
also excite density waves due to shear (Bodo et al. 2005;
Mamatsashvili & Chagelishvili 2007; Mamatsashvili & Rice
2009; Heinemann & Papaloizou 2009). In the 3D case, there
are a larger number of modes in a disc and it is quite possible
that some of them may appear linearly coupled due to shear
and, therefore, be able to generate each other during evo-
lution, especially the f-mode and the basic branch of the r-
mode (because they vary on comparable vertical scales in the
presence of self-gravity, Figs. 9, 10). Another related prob-
lem also of interest is the interaction between self-gravity
and the MRI in magnetized discs (see e.g., Fromang 2005).
In particular, how self-gravity can modify the growth rates
of magnetic normal modes responsible for the MRI. Actu-
ally, this will be the generalization of the extensive analysis
of normal modes in magnetized discs by Ogilvie (1998).
showed that
Lubow & Ogilvie (1998)
in non-self-
gravitating discs with polytropic vertical stratification, an
external forcing preferentially excites the f-mode, because it
has the largest responsiveness to an external driving com-
pared to other modes. This mode, propagating through a
disc, results in energy dissipation near the disc surface.
Based on these results and partly on the properties of f-
modes in stellar dynamics, Pickett et al. (2000) identified
the behaviour of 3D perturbations in self-gravitating discs
involving large surface distortions and the resulting energy
dissipation in the upper layers, with f-mode dynamics. The
self-stimulated potential was thought to play the role of an
external/tidal force. However, it is not obvious that the ef-
fect of a self-stimulated potential is the same as that of the
external potential. In fact, our analysis has revealed that in
self-gravitating discs, in addition to the f-mode, the r-mode
can also be dynamically important, because this mode ap-
pears to be subject to gravitational instability, while the
f-mode is not. The eigenfunctions of the gravitationally un-
stable basic even r-mode differ from those of the r-mode in
non-self-gravitating discs in that they are no longer concen-
trated near the midplane and behave somewhat similarly to
the eigenfunctions of the f-mode: they vary over the whole
vertical extent of the disc and also involve noticeable per-
turbations of the disc surface. Consequently, like the f-mode,
the gravitationally unstable r-mode can, in principle, also in-
duce gas motion causing large surface distortions and resul-
tant energy dissipation in the upper layers of the disc, which
is thought to play a role in enhancing disc cooling (because
the energy is deposited in the upper layers with small optical
depth, it can be radiated away more quickly and effectively
cool the disc, but this is a subject of further study, see e.g.,
Johnson & Gammie 2003; Boley et al. 2006). Furthermore,
in the case of non-axisymmetric perturbations, as mentioned
above, because of shear, the gravitationally unstable r-mode
can couple with and generate the strong f-mode. So, the
surface distortions may be caused by a combination of the
f- and r-modes. In order to explore where dissipation can
predominantly occur in a self-gravitating disc, one needs
to generalize the analysis of Lin et al. (1990), LP, LO98,
Bate et al. (2002) on the propagation of waves in stratified
non-self-gravitating discs and consider the propagation of
non-axisymmetric modes in stratified self-gravitating discs.
The dispersion properties of modes in the presence of self-
gravity as found here are one of the necessary things for
studying mode propagation.
Another point we want to raise concerns the spatial dis-
tribution of temperature. In order to realistically model the
cooling of protoplanetary discs, Boley et al. (2007) employed
the radiative transfer technique. In the vertical z−direction,
the radiative transfer equation was solved exactly assuming
a plane-parallel atmosphere approximation, but in the ra-
dial direction only the radiation diffusion approximation was
employed. However, as our linear results (Fig. 11) and other
non-linear simulations (e.g., Mej´ıa et al. 2005; Boley et al.
2006) demonstrate, temperature and, therefore, opacity may
vary on comparable scales in both the radial and vertical di-
rections and have very complex structure in the non-linear
regime. This implies that a more general radiative transfer
treatment based on solving the ray equation in all directions,
16 G. R. Mamatsashvili and W. K. M. Rice
rather than using the diffusion approximation in either di-
rection, would be more appropriate for better understanding
cooling processes.
ACKNOWLEDGMENTS
G.R.M. would like to acknowledge the financial support
from the Scottish Universities Physics Alliance (SUPA). We
thank Gordon Ogilvie for helpful discussions and for criti-
cally reading the manuscript. We also thank the referee for
the constructive comments that improved the presentation
of our results.
REFERENCES
Nelson A. F., 2006, MNRAS, 373, 1039
Ogilvie G. I., 1998, MNRAS, 297, 291
Papaloizou J. C., Savonije G. J., 1991, MNRAS, 248, 353
Papaloizou J. C. B., Lin D. N. C., 1989, ApJ, 344, 645
Pickett B. K., Cassen P., Durisen R. H., Link R., 2000,
ApJ, 529, 1034
Pickett B. K., Mej´ıa A. C., Durisen R. H., Cassen P. M.,
Berry D. K., Link R. P., 2003, ApJ, 590, 1060
Rafikov R. R., 2007, ApJ, 662, 642
Rice W. K. M., Armitage P. J., Bate M. R., Bonnell I. A.,
2003, MNRAS, 339, 1025
Rice W. K. M., Lodato G., Armitage P. J., 2005, MNRAS,
364, L56
Romeo A. B., 1992, MNRAS, 256, 307
Romeo A. B., 1994, A&A, 286, 799
Ruden S. P., Papaloizou J. C. B., Lin D. N. C., 1988, ApJ,
329, 739
Adams F. C., Ruden S. P., Shu F. H., 1989, ApJ, 347, 959
Bate M. R., Ogilvie G. I., Lubow S. H., Pringle J. E., 2002,
MNRAS, 332, 575
Bertin G., Lin C. C., Lowe S. A., Thurstans R. P., 1989,
Shu F., 1968, PhD thesis, Harvard University, Cambridge,
Massachusetts, USA
Stamatellos D., Whitworth A. P., 2009, MNRAS, 400, 1563
Tevzadze A. G., Chagelishvili G. D., Zahn J., 2008, A&A,
478, 9
Toomre A., 1964, ApJ, 139, 1217
Toomre A., 1981, in S. M. Fall & D. Lynden-Bell ed., Struc-
ture and Evolution of Normal Galaxies. Cambridge Uni-
versity Press, Cambridge
ApJ, 338, 104
Binney J., Tremaine S., 1987, Galactic dynamics. Princeton
University Press, Princeton, NJ
Bodo G., Chagelishvili G., Murante G., Tevzadze A., Rossi
P., Ferrari A., 2005, A&A, 437, 9
Boley A. C., 2009, ApJ, 695, L53
Boley A. C., Durisen R. H., 2006, ApJ, 641, 534
Boley A. C., Durisen R. H., Nordlund A., Lord J., 2007,
ApJ, 665, 1254
Boley A. C., Mej´ıa A. C., Durisen R. H., Cai K., Pickett
M. K., D'Alessio P., 2006, ApJ, 651, 517
Boss A. P., 1998, ApJ, 503, 923
Boss A. P., 2003, ApJ, 599, 577
Boss A. P., 2004, ApJ, 610, 456
Fromang S., 2005, A&A, 441, 1
Gammie C. F., 2001, ApJ, 553, 174
Goldreich P., Lynden-Bell D., 1965a, MNRAS, 130, 97
Goldreich P., Lynden-Bell D., 1965b, MNRAS, 130, 125
Goldreich P., Tremaine S., 1978, ApJ, 222, 850
Heinemann T., Papaloizou J. C. B., 2009, MNRAS, 397,
52
Johnson B. M., Gammie C. F., 2003, ApJ, 597, 131
Korycansky D. G., Pringle J. E., 1995, MNRAS, 272, 618
Laughlin G., Bodenheimer P., 1994, ApJ, 436, 335
Laughlin G., Korchagin V., Adams F. C., 1997, ApJ, 477,
410
Lin D. N. C., Papaloizou J. C. B., Savonije G. J., 1990,
ApJ, 364, 326
Lodato G., Rice W. K. M., 2004, MNRAS, 351, 630
Lodato G., Rice W. K. M., 2005, MNRAS, 358, 1489
Lubow S. H., Ogilvie G. I., 1998, ApJ, 504, 983
Lubow S. H., Pringle J. E., 1993a, MNRAS, 263, 701
Lubow S. H., Pringle J. E., 1993b, ApJ, 409, 360
Mamatsashvili G. R., Chagelishvili G. D., 2007, MNRAS,
381, 809
Mamatsashvili G. R., Rice W. K. M., 2009, MNRAS, 394,
2153
Mayer L., Lufkin G., Quinn T., Wadsley J., 2007, ApJ, 661,
L77
Mej´ıa A. C., Durisen R. H., Pickett M. K., Cai K., 2005,
ApJ, 619, 1098
|
1604.00859 | 1 | 1604 | 2016-04-04T13:53:03 | High Contrast Imaging with Spitzer : Constraining the Frequency of Giant Planets out to 1000 AU separations | [
"astro-ph.EP"
] | We report results of a re-analysis of archival Spitzer IRAC direct imaging surveys encompassing a variety of nearby stars. Our sample is generated from the combined observations of 73 young stars (median age, distance, spectral type = 85 Myr, 23.3 pc, G5) and 48 known exoplanet host stars with unconstrained ages (median distance, spectral type = 22.6 pc, G5). While the small size of Spitzer provides a lower resolution than 8m-class AO-assisted ground based telescopes, which have been used for constraining the frequency of 0.5 - 13 $M_J$ planets at separations of $10 - 10^2$ AU, its exquisite infrared sensitivity provides the ability to place unmatched constraints on the planetary populations at wider separations. Here we apply sophisticated high-contrast techniques to our sample in order to remove the stellar PSF and open up sensitivity to planetary mass companions down to 5\arcsec\ separations. This enables sensitivity to 0.5 - 13 $M_J$ planets at physical separations on the order of $10^2 - 10^3$ AU , allowing us to probe a parameter space which has not previously been systematically explored to any similar degree of sensitivity. Based on a colour and proper motion analysis we do not record any planetary detections. Exploiting this enhanced survey sensitivity, employing Monte Carlo simulations with a Bayesian approach, and assuming a mass distribution of $dn/dm \propto m^{-1.31}$, we constrain (at 95% confidence) a population of 0.5 - 13 $M_J$ planets at separations of 100 - 1000 AU with an upper frequency limit of 9%. | astro-ph.EP | astro-ph |
High Contrast Imaging with Spitzer : Constraining the Frequency of Giant
Planets out to 1000 AU separations.
Stephen Durkan1, Markus Janson2, Joseph C. Carson3
ABSTRACT
We report results of a re-analysis of archival Spitzer IRAC direct imaging surveys encompassing a va-
riety of nearby stars. Our sample is generated from the combined observations of 73 young stars (median
age, distance, spectral type = 85 Myr, 23.3 pc, G5) and 48 known exoplanet host stars with unconstrained
ages (median distance, spectral type = 22.6 pc, G5). While the small size of Spitzer provides a lower
resolution than 8m-class AO-assisted ground based telescopes, which have been used for constraining the
frequency of 0.5 - 13 MJ planets at separations of 10 − 102 AU, its exquisite infrared sensitivity provides
the ability to place unmatched constraints on the planetary populations at wider separations. Here we
apply sophisticated high-contrast techniques to our sample in order to remove the stellar PSF and open
up sensitivity to planetary mass companions down to 5′′ separations. This enables sensitivity to 0.5 - 13
MJ planets at physical separations on the order of 102 − 103 AU , allowing us to probe a parameter space
which has not previously been systematically explored to any similar degree of sensitivity. Based on a
colour and proper motion analysis we do not record any planetary detections. Exploiting this enhanced
survey sensitivity, employing Monte Carlo simulations with a Bayesian approach, and assuming a mass
distribution of dn/dm ∝ m−1.31, we constrain (at 95% confidence) a population of 0.5 - 13 MJ planets
at separations of 100 - 1000 AU with an upper frequency limit of 9%.
Subject headings: planetary systems - techniques: image processing - infrared: planetary systems
1.
Introduction
The plethora of confirmed exoplanets to date is
dominated by a population of planets within 5 -- 6
AU of their host star. This has been due to the
success of
large-scale radial-velocity (Butler et al.
1996; Vogt et al. 2000; Mayor et al. 2003) and tran-
sit surveys (Pollacco et al. 2006; Baglin et al. 2009;
Koch et al. 2010), which are biased to short separa-
tion planets and whose typical duration limit their
ability to detect periodicities on the order of 10's of
years. Such surveys account for a combined detec-
tion of ∼ 1900 planets (See http://exoplanet.eu and
http://exoplanetarchive.ipac.caltech.edu/), > 90% of
1Astrophysics Research Centre,
of Mathemat-
ics & Physics, Queen's University, Belfast BT7 1NN, UK;
[email protected]
School
2Department of Astronomy, Stockholm University, 106 91
Stockholm, Sweden
3Department of Physics and Astronomy, College of Charleston,
Charleston, SC 29424 , USA
1
the entire exoplanet population. This sample enables
statistically significant trends to be uncovered between
planetary / host star properties and planet frequency,
which in turn allows for planet formation and evolu-
tion theories to be stringently tested and constrained.
However formation scenarios and evolutionary paths
for wide giant planets with separations >> 5 -- 6 AU
continue to prove challenging to constrain. This is due
to a lack of systematic explorations by surveys with a
sufficient degree of sensitivity and statistically robust
population analyses.
Direct imaging provides the most viable technique
to probe for giant planets at such wide separations. Ex-
tensive work has been carried out to expand the sam-
ple of wide giant planets through detection in large-
scale direct imaging surveys (Masciadri et al. 2005;
Biller et al. 2007; Lafreni`ere et al. 2007b; Kasper et al.
2007; Chauvin et al. 2010; Leconte et al. 2010; Ehrenreich et al.
2010; Carson et al. 2011; Janson et al. 2011; Vigan et al.
2012; Delorme et al. 2012; Biller et al. 2013; Chauvin et al.
2015). Whilst these surveys typically record non-
detections, several planetary mass wide companions
have been found, indicating that although these types
of planets may be rare (Nielsen & Close 2010), they do
exist with some frequency throughout the galaxy. By
subjecting these imaging surveys to statistical analysis,
an upper limit on this frequency can be determined.
Concurring frequency upper limits have been found
for wide giant planets over a separation range on the
order of 101 - 102 AU, corresponding to the parame-
ter space at which observations are sensitive to Jupiter
mass companions. These limits are shown in Table 1.
Sensitivity is confined to this range due to instru-
mental limitations. Imaging surveys typically favour
the use of adaptive optics (AO) corrected instruments
on 8m class ground based telescopes. The high angu-
lar resolution afforded by such instruments provides
sensitivity to planetary mass companions at small
separations, down to the order of 10 AU, inside of
which the required contrast becomes unachievable as
one approaches the core of the near diffraction lim-
ited point spread function (PSF). The outer sensitivity
limit stems from anisoplanatism, where AO delivers
poor wave front correction at increasing separation
due to the different propagation paths and hence vary-
ing wave front distortion experienced by off-axis light.
This substandard correction results in a decrease in
image quality and therefore a reduction in sensitivity
at large separations. Typical values for the isoplanatic
angle, where distortion is statistically uniform allow-
ing for good AO correction, are 10′′ - 20′′ in the near
infrared (NIR) (Sandler et al. 1994; Fritz et al. 2010),
and thus imaging instruments are typically restricted
to this field of view (FOV). This has severely limited
sensitivity to planetary mass companions at separa-
tions beyond the order of 102 AU, for typical nearby
targets.
Conducting imaging surveys from space based tele-
scopes would negate this effect, however their small
aperture diameters produce large diffraction limited
PSF's, severely limiting their application for direct
imaging of planets. Still, Marengo et al. (2006, 2009)
have shown that the Spitzer space telescope is ca-
pable of sensitivity to planetary mass companions at
large angular separations, within the background noise
limited regime, with their studies of ǫ Eri and Fo-
mauhaut. Recent studies of Vega, Fomalhaut and ǫ Eri
by Janson et al. (2012, 2015) have demonstrated that
sensitivity to planetary mass companions is achievable
with Spitzer within the PSF noise-limited regime, with
the application of sophisticated high-contrast reduc-
tion techniques.
Therefore we implemented a sophisticated PSF
subtraction technique to enhance the sensitivity of
archival Spitzer imaging surveys, enabling sensitiv-
ity to planetary mass companions over a separation on
the order of 102 - 103 AU. This parameter space has
not previously been systematically explored by sur-
veys to a sufficient degree of sensitivity, leaving the
population of giant planets at these separations poorly
constrained.
While the formation and origin of such wide
orbit giants proves difficult to explain,
their exis-
tence has been confirmed with the detection of sev-
eral planetary mass companions e.g. 1RXS J1609-
2105 b; 330 AU (Lafreni`ere et al. 2010), FW Tau b;
330 AU (Kraus et al. 2014a), HD106906 b; 650 AU
(Bailey et al. 2014), GU Psc b; 2000 AU (Naud et al.
2014). Whilst core accretion and gravitational instabil-
ity modes cannot readily explain the formation of such
planets in situ (Ida & Lin 2004; Boss 2006; Rafikov
2007; Dodson-Robinson et al. 2009), and disk interac-
tions cannot viably migrate planets out to ∼ 103 AU,
beyond the typical confines of the disk (Isella et al.
2009), several theories have been developed in an at-
tempt to account for planetary existence at these large
orbital separations.
One suggestion is that such a planetary system
could be the result of dynamical capture of free-
floating planets during dispersal of a stellar cluster.
Simulations found this capable of producing planets
at separations of > 50 AU (Parker & Quanz 2012)
and 102 - 105 AU (Perets & Kouwenhoven 2012).
As an alternative, planet-planet scattering, in which
multiple planets gravitationally interact after disk
dissipation, can result in dynamical scattering of a
planet out to wide separations on the order of 100's
of AU (Rasio & Ford 1996; Veras & Armitage 2004;
Chatterjee et al. 2008; Juri´c & Tremaine 2008). Dy-
namical simulations by Veras et al. (2009) show these
interactions capable of scattering giant planets out to
separations of 102 - 105 AU. However they find the
population of planets that eventually end up in un-
bound orbits passing through these wide separations
to be larger than the population on stable orbits. The
total population of wide giant planets then decreases
on timescales of ∼ 10 Myr to produce a significantly
depleted population at ages > 50 Myr where the ma-
jority of planets have been ejected from the system.
Constraining the population of giant planets at sep-
arations of 102 - 103 AU is essential for assessing the
2
Table 1: Planet Frequency Upper Limits
Planet Frequency
Mass Range (MJ )
Separation Range (AU)
Fractional Upper Limit
Study
0.5 - 13.0
0.5 - 13.0
1.0 - 20.0
1.0 - 13.0
50 - 250
25 - 100
10 - 150
20 - 150
0.093
0.110
0.060
0.100
(Lafreni`ere et al. 2007b)
(Lafreni`ere et al. 2007b)
(Biller et al. 2013)
(Chauvin et al. 2010)
relevance of these theories, and allowing for stringent
constraints to be placed on formation and evolution
modes out to 1000's of AU.
2. Target Sample
Our target list is compiled from archival Spitzer
data based on proximity and youth. Such targets prove
favourable to imaging searches for planetary compan-
ions as young giant planets are intrinsically bright in
the near infrared due to heat retention from forma-
tion. This brightness decreases as the planet ages and
the heat dissipates (Baraffe et al. 2003; Burrows et al.
2003; Fortney et al. 2008). Younger targets therefore
provide imaging sensitivity to a greater range of com-
panion masses for a given detection threshold, com-
pared to relatively older stars, allowing for detection
of lower mass planets. Stars within close proxim-
ity to Earth provide sensitivity to companions over a
greater range of physical separations for a given de-
tection threshold at a specific angular separation, com-
pared to a relatively distant target, allowing for detec-
tion of shorter period planets for a given mass sensitiv-
ity.
This motivates our choice of stars from the archival
Spitzer program 34 (P34) for enhanced imaging anal-
ysis. This program targeted 73 nearby (< 30 pc),
young stars and therefore provides an ideal basis for
the sample of targets chosen here. Multiple age in-
dicators including X-ray luminosity, chromospheric
activity, Lithium abundance, rotation and photometric
colour were originally used to select targets for P34
with ages below 120 Myr. We find these age estimates
to be overly optimistic. Using a number of sources
(e.g. Montes et al. 2001b; Zuckerman & Song 2004;
Torres et al. 2008; Maldonado et al. 2010; Malo et al.
2013), we identify 55 of these targets to be members of
young moving groups (YMGs) and place conservative
age limits on the targets corresponding to reliable age
estimates of YMGs taken from the literature; these age
estimates are given in Table 2. We then place conser-
vative age limits on the remaining targets combining
literature age estimates encompassing several tech-
niques (e.g. Barnes 2007; Mamajek & Hillenbrand
2008; Plavchan et al. 2009; Tetzlaff et al. 2011; Vican
2012). The median target age is 85 Myr, although age
limits range from 8 to 1050 Myr. The exception to
this is HD 124498. Malo et al. (2013) identifies this
target as a probable member of the β Pictoris YMG.
However Chauvin et al. (2010) use several age dating
techniques to reclassify HD 124498 as an older sys-
tem with an age ≥ 100 Myr. We therefore place an
age of 12 Myr - 10 Gyr on HD 124498 to ensure any
statistics encompassing age estimations produce con-
servative results. The median target distance, spectral
type and H band magnitude are 23.3 pc, G5 and 5.289
respectively.
We choose an additional archival Spitzer program
to add to our target sample, program 48 (P48). This
program targeted 48 nearby (< 35 pc) stars with
known planetary companions, discovered via radial
velocity. As such, these planets are on relatively short
orbits, spanning a parameter space currently inacces-
sible to direct imaging techniques. These exoplanet
host stars are relatively old and no confident age limits
can be found for the majority of the sample through-
out the literature. For consistency, and to justify that
any derived statistical result represents a conservative
limit, we adopt a conservative age of 1 Gyr - 10 Gyr
for all P48 stars. This age range spans the breadth of
poorly constrained literature ages for the majority of
the sample. The exceptions to this are HD 13507, HD
1237 and AF Hor, which are identified as members of
YMGs with independent literature age estimates that
support the association membership. Whilst these tar-
gets may prove unfavourable to a deep imaging study,
due to their extended ages, the strength in their addi-
tion to the sample is that they provide an additional 48
references to aid in PSF reduction of the 73 P34 stars,
contributing to an increase in achievable contrast and
sensitivity to smaller mass companions. P48 was exe-
cuted under the exact observational parameters as P34.
3
Table 2: Moving Group Age Estimates
YMG
Age Estimate (Myr)
Age Reference
Local Association (LA)
β Pictoris
Ursa Major
Castor
IC2391
Hyades
Her-Lyr
AB Dor
Octans-Near
Argus
TW Hydrae
Tuc-Hor/Columba/Carina
20 - 150
12 - 22
400 - 600
100 - 300
45 - 55
575 - 675
211- 303
70 - 120
30 - 100
30 - 50
8 - 12
20 - 40
(Montes et al. 2001b; Brandt et al. 2014)
(Malo et al. 2013)
(King et al. 2003)
(Barrado y Navascues 1998; Montes et al. 2001b)
(Stauffer et al. 1997)
(Perryman et al. 1998)
(Eisenbeiss et al. 2013)
(Malo et al. 2013)
(Zuckerman et al. 2013)
(Malo et al. 2013)
(Malo et al. 2013)
(Malo et al. 2013)
Along with P48 encompassing a similar spectral sam-
ple of stars as P34, median spectral type and H band
magnitude G5 and 4.957 respectively, this ensures P48
targets provide sufficient P34 PSF references. The me-
dian distance of P48 stars is 22.6 pc. So whilst the age
range of P48 stars limit sensitivity to low mass plan-
ets, their proximity ensures sensitivity to larger mass
planets over a wide range of separations. Therefore
their inclusion our sample for reduction and analy-
sis as well as providing additional references is jus-
tified. However we note the potential bias introduced
to any population constraint statistically derived from
a deep imaging search encompassing a sample of stars
hosting short period planets.
If we consider planet-
planet scattering to be a relevant mechanism for gi-
ant planet production at 102 - 103 AU separations, the
presence of a detected planet at a separation compara-
ble to where a wide giant planet is expected to initially
form, incorporates some degree of bias into our statis-
tics that we cannot quantitatively evaluate. Therefore
we do not correct for this possible bias. The combined
target properties are given in Table 3.
3. Observations and Data Reduction
The combined 121 targets were observed with the
Infrared Array Camera (IRAC; Fazio et al. 2004) on
the Spitzer Space telescope (Werner et al. 2004) under
archival programs 34 and 48 between 2003 and 2004,
during the cryogenic phase of the Spitzer mission. Im-
ages were obtained simultaneously at 3.6, 4.5, 5.8 and
8 µm (channels 1 - 4 respectively). We choose Channel
2 as the primary channel to analyse based on evolution-
ary model predictions that planetary mass companions
are at their peak luminosity within the 4.5 µm band
(Barrafe et al. 2003). Therefore this channel offers
the best sensitivity for planet detection. We also anal-
yse IRAC channel 1 images as a means to vet potential
companions.
All targets were observed with a 30s frame time
allowing for an effective exposure time of 26.8s per
dither position. Each observation was carried out
with a five-position Gaussian dither pattern. The
Spitzer Science Center (SSC) IRAC Pipeline (version
S18.25.0) performed data reduction for all observa-
tions, producing basic calibrated data (BCD) frames.
This pipeline also produced corrected BCDs (CBCD)
where saturated point sources have been fitted with un-
saturated point sources by fitting an appropriate PSF
that is matched to the unsaturated wings of the source.
A final post basic calibrated data (PBCD) frame is
then created by mosaicking the relevant CBCD frames
at each dither position to produce a sub-pixelated im-
age with a 0.6′′ per pixel resolution and a 5.2′ x 5.2′
FOV. Such saturation corrected images are unsuitable
for a close companion search as pixels towards the
PSF core may have been replaced with model pixel
values, ensuring that any information about potential
companion sources at these separations is lost. How-
ever we favour the use of PBCD's over BCD's due to
the additional artifact correction performed on the for-
mer by the SSC IRAC pipeline. Therefore we generate
a composite frame consisting of a PBCD image where
saturation corrected regions, identified in the relevant
mask files, have been replaced with pixel values gener-
ated from mosaicking the relevant BCD frames. These
pixel values have been sub-pixelated to the equivalent
0.6′′ per pixel resolution. Figure 1 shows an example
4
Fig. 1. -- Example of a composite PBCD / BCD 4.5 µm image of HD 217813, displaying the extent of the PSF and
Spitzer spider features. The wide FOV reveals a multitude of point sources which must be vetted for planet candidacy.
HD 217813 properties lie close to the median of the sample, H mag = 5.232, G5V star at 24.7pc.
Fig. 2. -- Final reduced image of HD 217813 depicting the same FOV as Figure 1. The white box highlights the 2.01′
x 2.01′ PCA optimization region. Regular PSF subtraction of a mean stack image has been performed outside the
optimisation region.
5
of such a composite PBCD / BCD image.
3.1. PSF Subtraction
We calculate the centre of each target PSF by the
fitting of a 2 dimensional Gaussian, the centre of each
image is then shifted to this position. Each image is
then normalised by the central brightness of the stellar
PSF, estimated from the saturation corrected images.
This provides an image stack of target PSF's that have
similar normalised flux values. We then mask the satu-
rated cores of each image for subsequent image reduc-
tion. Each individual image is then robustly centred
with respect to the stack by subtraction of a median
stack image over a range of sub-pixel offsets. The off-
set producing the minimum residuals is then used to
align each image. The image stack constitutes the li-
brary of PSF's used to construct a reference PSF for
image subtraction. This accurate alignment of target
PSF's then provides better performance of any refer-
ence construction, ultimately leading to improved sen-
sitivity in any final reduced image.
Here we use Principal Component Analysis (PCA)
to construct the optimal reference PSF to subtract from
the target PSF. Variations of PCA such as PynPoint
(Amara & Quanz 2012) and KLIP (Soummer et al.
2012) have the same underlying principle; a linear
combination of orthogonal basis sets are used for ref-
erence construction. These basis sets represent the de-
composition of the reference library into its principal
components. The linear coefficients are then gener-
ated by the projection of the target onto each individ-
ual basis. PCA performed here follows a KLIP-based
analysis.
We limit the PCA optimization to a 201 x 201 pixel
sub-section of each 3.6 µm and 4.5 µm image, centred
on the star. This corresponds to 2.01′ x 2.01′ FOV.
This reduced area is chosen as a favourable trade off
between algorithm efficiency and sensitivity to wide
separations, with 2.01′ corresponding to separations
on the order of 103 AU, at the typical target distance.
PCA is performed on concentric annuli centred on
the star. The radii of the annuli are chosen such that
each annulus contains 1500 pixels. Reference annuli
containing astronomical sources such as background
stars, stellar companions or surviving bad pixels are
excluded to prevent the algorithm from subtracting any
true planetary PSF , and to prevent any fake planetary
signal being injected into the final image. Since flux
increases towards the PSF core, the larger flux val-
ues towards the inner region of each annulus dominate
the weighting of the orthogonal basis. This may pro-
duce the lowest residuals but the outer annular regions
will have experienced a substandard reduction. By per-
forming PCA four times on each target and increasing
the radius of the inner saturation mask, and therefore
shifting the radii of the 1500 pixel annuli, we ensure
that each section of the frame has been encompassed
by an inner annular region and experienced a full qual-
ity reduction. The final images to be analysed are then
composites of these four separate PCA reduced frames
whose constituent parts have experienced optimal re-
duction. A final PCA reduced image can be seen in
Figure 2.
PCA reduction is typically performed at neighbour-
ing states of the telescope / instrument system and
provides optimal PSF subtraction when carried out in
tandem with angular differential imaging application,
where the PSF remains quasi-static over the observa-
tion. Here we apply PCA to a sample observed at vary-
ing states of the telescope / instrument system taken
over the course of one year, e.g. varying stellar mag-
nitude, spectral type, telescope roll angle. Therefore,
our study constitutes a trial of PCA on an unfavourable
data set, testing its effectiveness at the limits of its ap-
plication. Similar conditions are required for optimal
PSF subtraction using LOCI (Lafreni`ere et al. 2007a)
and thus we favour PCA for its speed enhancement.
3.2.
Image Sensitivity & Companion Identifica-
tion
We evaluate the sensitivity in the final PCA reduced
images at 5σ detection limits by relating the standard
deviation in concentric 1 pixel width annuli centred
on the star, to the zero-point flux of Vega using the
method of Marengo et al. (2009). This generates sen-
sitivity values in units of Vega magnitudes. As any po-
tential companion will suffer from partial flux subtrac-
tion during the PCA reduction, we must account for
the throughput of sources in the sensitivity estimation.
We inject synthetic companions into each target im-
age at position angles 0◦ and 90◦ so as to cover spider
and non-spider regions, at five pixel intervals out to the
edge of the optimisation region, and estimate through-
put after PCA application. The mean throughput is
then calculated at each separation and linearly interpo-
lated to produce throughput values for each pixel sep-
aration, consistent with sensitivity estimation in one-
pixel width annuli. The mean throughput over the im-
age stack is then used for sensitivity estimation.
6
Fig. 3. -- PCA sensitivity improvement with respect to conventional PSF subtraction, as a function of separation.
The decrease in improvement around 20′′ corresponds to the transition between the PSF noise-limited regime and the
background noise limited regime. The jagged nature of the curve stems from preferential reduction of inner annulus
regions.
7
We evaluate the performance of the PCA reduction
by computing sensitivity curves for both PCA opti-
mized and conventional PSF subtracted images (i.e.,
subtraction of a mean stack image) and calculating
the sensitivity improvement provided by PCA for each
target. The median sensitivity improvement curve is
then generated. This curve, displayed in Figure 3,
shows that PCA offers superior image sensitivity at
separations less than 20′′, within the PSF noise-limited
regime, and provides a median improvement of ∼ 0.9
magnitudes at separations less than 10′′. This validates
the application of PCA in a study encompassing a di-
verse sample of stars, observed at varying states of the
telescope / instrument system. The ∼ 0.9 magnitude
improvement in the contrast limited regime demon-
strates the relevance of applying and developing so-
phisticated high-contrast techniques to archival Spitzer
data, where we have enhanced sensitivity to planetary
mass companions at relatively small angular separa-
tions. This allows for the possibility of subsequent
planet detection within a previously elusive parame-
ter space and enabling more stringent constraints to
be placed on the wide giant population. In the back-
ground noise limited regime, separations greater than
20′′, PCA does not provide any sensitivity improve-
ment over regular PSF subtraction. This is due to the
random nature of background noise which PCA cannot
reference.
We use two initial criteria to identify potential plan-
etary mass companions. Firstly each 4.5 µm reduced
image is visually inspected to identify realistic PSF
shapes, allowing real sources to be distinguished from
potential surviving bad pixels. Any realistic PSF is
then vetted for planetary candidacy by comparison of
the 4.5 µm and 3.6 µm images. The spectral energy
distribution of a non-irradiated 0.5 - 13 Jupiter mass
companion, at the typical ages of our sample, is such
that planetary flux at 3.6µm is typically > 1 magni-
tude fainter than at 4.5 µm where peak emission oc-
curs (Baraffe et al. 2003; Spiegel & Burrows 2012). In
contrast to a background star, which is expected to be
approximately equally bright in the 3.6 µm and 4.5
µm images, a planet detection recorded around the 5σ
limit at 4.5 µm is not expected to be recovered to any
reasonable significance at 3.6µm. This allows plane-
tary sources to be distinguished from faint background
stars without the need for a proper motion analysis.
Therefore we look for source non-detection at 3.6 µm
to confirm planet candidacy.
4. Results
4.1. Observational Sensitivities
Figures 4 and 5 show the 4.5 µm median sensi-
tivity curves for the P34 and P48 stars respectively.
Within ∼ 20′′ sensitivity decreases towards the PSF
core where the residual PSF noise exhibits the largest
variance, limiting sensitivity to lower magnitude com-
panions. Outside ∼ 20′′ PCA provides no sensitivity
improvement and we are limited by background noise
which tends to be constant, thus magnitude sensitivity
at these separations is roughly constant. Figures 4 and
5 also map sensitivity as a function of projected physi-
cal separation out to ∼ 1400 AU at the median distance
of P34 and P48 stars, 23.3 pc and 22.6 pc respectively.
P34 and P48 stars provide comparative magnitude sen-
sitivity limits.
These magnitude sensitivity limits can be translated
into mass sensitivities using mass-luminosity evolu-
tionary models. One caveat is the discrepancy be-
tween hot- (Chabrier et al. 2000; Baraffe et al. 2003;
Burrows et al. 2003) and cold-start
(Marley et al.
2007; Fortney et al. 2008) models, where the later
predict much fainter planets at young ages. Here
we will consider hot-start models which are consis-
tent with existing observational data (Janson et al.
2011). However it can be noted that the models
converge on the order of 10's of Myrs for low mass
planetary companions (≤ 2MJ) and on the order of
100's of Myr for higher mass planetary companions
(Spiegel & Burrows 2012). Thus our choice of model
will not lead to significant disparity in mass sensi-
tivity over the complete sample at the typical ages
considered. COND-based models (Allard et al. 2001;
Baraffe et al. 2003), applicable for companion temper-
atures below 1700K, which is a relevant temperature
range for 0.5 - 13 MJ companions at the sample ages,
are used here.
The corresponding mass sensitivities for the median
P34 and P48 magnitude sensitivities are shown in Fig-
ures 4 and 5. With P34 targets, sensitivity down to
≤ 2MJ companions is achieved down to ∼5′′, cor-
responding to ∼ 100 AU projected separation, whilst
sensitivity down to 0.5 MJ companions is achieved for
separations & 15′′ (& 350 AU). Whilst P34 and P48
stars provide comparative magnitude sensitivity lim-
its, the extended ages of P48 stars ensure an equivalent
low mass sensitivity cannot be achieved. Sensitivity to
below 4 MJ is typically not acquired, and sensitivity
to 5 MJ planets is limited to outside 20′′ (∼ 500 AU).
8
Fig. 4. -- Median survey detection limits for P34 stars. Magnitude sensitivity is given as a function of angular
separation. Right axis shows corresponding minimum detectable mass at median target age of 85 Myr. Top axis
displays projected physical separation at median target distance 23.3 pc.
Fig. 5. -- Median survey detection limits for P48 stars. Right axis shows corresponding minimum detectable mass at
median target age of 5.5 Gyr. Top axis displays projected physical separation at median target distance 22.6 pc.
9
Taking 13 MJ as the upper mass limit for planetary
objects, we are typically not sensitive to planets within
10′′ (∼ 200 AU) for these targets.
4.2. Candidate Companion Detection
Through comparison of 3.6 µm and 4.5 µm im-
ages we initially identify 36 candidates with realis-
tic planetary mass colour. Figure 6 shows a typical
candidate identification through a 4.5 µm source non-
detection at 3.6 µm. These candidates are further vet-
ted with a common proper motion analysis using a 2nd
epoch combining archival Spitzer data and observa-
tions carried out during Spitzer Cycle 11 under pro-
gram 11102, repeating the original observations for
several candidate host stars under the same observa-
tional parameters, with subsequent equivalent reduc-
tion. The target sample has a median total proper
motion of 206 mas/yr, and over a baseline of 4 - 11
years provided by the 2nd epoch data, we can con-
fidently identify co-moving planetary companions in
an image with a 0.6′′ /pixel scaling. Four candidates
recovered in a 2nd epoch are revealed to be non co-
moving with the target star. Such sources are most
likely rare background galaxies with unusual infrared
colors, such as NGC 1377 which is brighter at 4.5 µm
than 3.6 µm (Dale et al. 2005). Non-detection of the
remaining sources and inspection of the raw CBCD
frames leads to their identification as bad pixels sur-
viving the Spitzer Science Center IRAC Pipeline re-
duction. Therefore this survey records a null planet
detection result.
5. Statistical Analysis
As in previous direct imaging surveys we exploit
this null detection and the magnitude detection limits
generated for each target to place constraints on the
wide giant population through a coupling of Monte
Carlo simulations and Bayesian analysis. Effectively
the Bayesian analysis determines the population of
wide giants, as an upper fractional limit of stars that
harbour such a companion, that is consistent with the
derived planet detection probability and the null sur-
vey result. We formulate our statistical analysis in the
same fashion as previous works based on Carson et al.
(2006) and Lafreni`ere et al. (2007b). The relevance
of this work is that with improved sensitivity in the
PSF-noise limited regime, due to PCA application to
archival Spitzer data, and the wide FOV providing
background noise limited sensitivity, we have opened
sensitivity to planetary mass companions over separa-
tions on the order of 102 - 103 AU. This has been done
for both young, P34, and relatively old, P48, stars, as
seen in Figures 4 and 5. With this enhanced sensitivity,
we can significantly constrain the population of wide
giants out to 1000 AU.
5.1. Detection Probabilities
To derive the planet detection probability for each
target we simulate 10,000 planets, using a Monte Carlo
approach to randomly sample planet mass, separa-
tion, orbital projection and age. Planet mass is sam-
pled between 0.5 - 13 MJ , however the lack of con-
straints on the planet population over the parameter
space explored here has ensured any constraints on
mass distribution are correspondingly lacking. There-
fore we assume a mass distribution of dn/dm ∝ mα
where α = −1.31, extrapolated from statistical anal-
ysis of radial velocity studies (Cumming et al. 2008).
In any case Chauvin et al. (2010) find the choice of
semi-major axis power law index to dominate the
derived detection probabilities in comparison to any
variation in α. We choose to sample semi-major
axis from a linear distribution between 100 - 1000
AU. Studies (Nielsen et al. 2008; Nielsen & Close
2010; Chauvin et al. 2010) have found semi-major
axis power laws reported by Cumming et al. (2008)
to be invalid at the separations considered here, mo-
tivating our choice of a linear distribution. We then
sample orbital projection factor using the method of
Brandeker et al. (2006) which accounts for orbital
phases, orientations and an eccentricity distribution
of f (e) = 2e (predicted and observed for long period
binaries; Duquennoy & Mayor 1991), allowing true
physical separations to be translated to projected phys-
ical separations. Projected separations are then con-
verted to angular separations using the known stellar
distance. Here we do not sample over any uncertainty
in stellar distance as the revised Hipparcos measure-
ments have good precision and thus the uncertainty
in stellar age completely dominates the uncertainty in
detection probabilities. Age is sampled between limits
of reliable estimates from the literature, or between 1
and 10 Gyr for the majority of P48 targets with ages
that are not well constrained. These ages are given in
Table 3. Applying COND-based evolutionary mod-
els and adopting sampled planetary properties, we can
translate mass into magnitude and map each planet
onto 5σ magnitude sensitivity curves. The fraction
of planets that lie above the detection limit provides
10
Fig. 6. -- PCA reduced 3.6 µm, left, and 4.5 µm, right, images of HD 197890. Majority of point sources revealed
appear at both wavelengths and therefore are likely background stars. Sources highlighted in green circles however
only appear at 4.5 µm and are identified as potential planetary candidates. After further analysis these apparent sources
were revealed to be surviving outlier pixel values.
the detection probability for each target. Mean Detec-
tion probabilities for P34 and P48 stars for mass range
[0.5, 13] MJ and the separation range [100, 1000] AU
are 0.42 and 0.08 respectively.
We additionally choose to perform simulations over
varying mass and separation ranges in order to con-
strain the wide giant population as a function of mass
and separation. Mass range will be varied between
[0.5, mmax] MJ with mmax ranging from 1.0 - 13.0
MJ in increments of 0.5 MJ . Separation range will
be varied between [amin, amax] with amin = 75, 100,
125 AU, and amax increasing in increments of 25 AU
out to 1000 AU. Mean detection probabilities over the
entire sample, as a function of upper mass and outer
separation limit, are given in Figure 7. Mean detec-
tion probabilities over P34 targets are typically much
greater than the mean over the entire sample due to the
superior mass sensitivity of P34 targets corresponding
to their relatively young ages. However we find the in-
clusion of P48 targets provides a favourable trade off
between low probability detection and increased sam-
ple size which leads to better statistical constraints on
the wide giant population.
5.2. Binary Bias
We identify 43 targets in our sample as binary stars
in the Washington double star catalogue (Mason et al.
2001) that have been vetted for true companionship
through a proper motion analysis. The inclusion of
these binaries, > 1/3 of the sample, may introduce a
bias in astrophysical interpretation which we attempt
to account for. We recored a null planet detection in
these binary systems, which is the basis of our sta-
tistical formalism. However the binary companion
will have introduced a parameter space of instabil-
ity in which we would not expect a planet to orbit
around its host star. To deal with this and remove the
bias from our analysis we use the stability criteria of
Holman & Wiegert (1999) to determine the instability
regions, for both S- and P-type systems. Many of the
binary orbits and companion masses are not well doc-
umented, so for algorithm simplicity, consistency and
to ensure conservative results, we take the worse-case
scenario of an equal mass binary system, which intro-
duces the greatest range of instability. We also assume
the latest epoch separation in the Washington double
star catalogue to be the true physical separation. This
is a reasonable assumption as for a random distribu-
11
Fig. 7. -- Mean planet detection probabilities over P34 and P48 targets as a function of semi-major axis and mass.
X axis denotes outer limit amax over which planet separation is sampled with constant amin = 100 AU. Curves are
labelled with upper mass limit mmax over which planet mass is sampled in units of MJ , with constant mmin = 0.5MJ.
12
tion of binary eccentricities, phases and orientations
the most likely true physical separation is given by
the projected separation (Brandeker et al. 2006). We
ensure these instability regions are counted as a non-
detectable range. We note that the true frequency of
wide giant planets around binary stars may be differ-
ent than that around single stars. This potential fre-
quency discrepancy is possibly due to the influence
that a binary companion will have on the planet for-
mation process. Whilst studies have found that bina-
rity has a minimal effect on overall planet frequency
(Bonavita & Desidera 2007; Bergfors et al. 2013), it
has been suggested that binary companions with sep-
arations . 100 AU may result in a decrease in the
number of planets formed through enhanced dynami-
cal heating of the protoplanetary disk (Thalmann et al.
2014). Whilst binary companions with separations
. 100 AU only account for 12 of the 43 binaries in
our sample, an element of bias will be inherent in the
statistical result derived from a combined single and
binary star sample.
5.3. Estimation of Planet Frequency
With the determination of target detection probabil-
ities an upper limit on the frequency of planets hosting
wide giants can be derived through the Bayesian ap-
proximation;
fmax ≈ −ln(1 − α)/N hpj i
(1)
Where hpji is the mean detection probability over
N targets and α is credibility level of the derived result
which we choose to be 95%. We derive an upper limit
on the frequency of planets in the mass range [0.5, 13]
MJ and the semi-major axis range [100, 1000] AU to
be 9%. As no previous survey has probed this param-
eter space to a similar degree of sensitivity there is
no literature comparison to our result. However our
low frequency findings certainly support the extension
of previous survey findings of low frequency at sep-
arations on the on the order or 10 - 100 AU (see Ta-
ble 1), out to 1000 AU separations, and is in general
agreement with the theory of Veras et al. (2009), pre-
dicting a low frequency population of 102 − 105 AU
planets at ages > 50 Gyr. Figure 8 shows planet fre-
quency limits as a function of mass over constant sep-
aration range 100 - 1000 AU. Figure 9 shows planet
frequency as a function of separation with amin = 75,
100 and 125 AU over constant mass range 0.5 - 13
MJ . Whilst this study well constrains the population
13
of 100 - 1000 AU planets with masses ranging from 0.5
- 13 MJ , the planet frequency upper limit increases to-
wards smaller mass and separation limits. This is a de-
terioration of the population constraint due to bias in-
herent in the detection technique, where imaging sen-
sitivity, and therefore detection probability, decreases
with decreasing mass and separation. However, planet
frequency tends to be a smoothly varying function with
respect to mass and separation, and Figure 9 shows
that at separations & 700 AU frequency values tend to
be approximately constant, corresponding to the back-
ground noise limited regime with approximately con-
stant sensitivity. Thus there is no reason to expect that
our choice of [0.5, 13] MJ and [100, 1000] AU bound-
aries are arbitrarily optimistic with regard to reason-
ably lower mass and separation limits.
6. Conclusions
In this paper we have presented the results of a re-
analysis of two archival Spitzer imaging surveys en-
compassing 121 targets with varying spectral types and
ages. Previously, the large PSF associated with the
0.85m Spitzer telescope diameter has severely limited
its capability for directly imaging exoplanets. With
the application of PCA we have removed the stel-
lar PSF and opened up sensitivity to planetary mass
companions over a broad range of separations. PCA
has provided up to a magnitude sensitivity improve-
ment at small separations with respect to conventional
PSF subtraction methods, highlighting the strength of
the technique, even on a relatively unfavourable data
set. Using theoretical mass-luminosity evolutionary
models we have shown that we are sensitive to plan-
etary mass companions down to 0.5 MJ at separa-
tions on the order of 102 − 103 AU. This parameter
space has not previously been systematically explored
by imaging surveys to any comparable degree of sen-
sitivity due to anisoplanatism and FOV limitations of
ground based surveys, and PSF contrast limitations of
space based surveys. Therefore through the coupling
of Monte Carlo simulations and a Bayesian analysis,
for the first time we have constrained the population
of 0.5 - 13 MJ , 100 - 1000 AU planets, producing an
upper frequency limit of 9%. This is an extension of
findings of low companion frequencies in numerous
previous surveys at separations on the order of 10−102
AU. Constraining this very wide giant planet popula-
tion allows for previously untested formation and evo-
lutionary theories to be adapted and constrained.
Fig. 8. -- Planet frequency upper limit (at the 95% confidence level) as a function of mass. X axis denotes upper mass
limit mmax with constant mmin = 0.5MJ. Separation range is constant, 100 - 1000 AU.
Fig. 9. -- Planet frequency upper limit (at the 95% confidence level) as a function of separation. X axis denotes outer
separation limit amax with constant amin = 75, 100, 125 AU. Mass range is constant, 0.5 - 13 MJ .
14
S.D. acknowledges support from the Queen's Uni-
versity Belfast Department for Education and Learn-
ing (DEL) university scholarship. M.J. gratefully ac-
knowledges funding from the Knut and Alice Wallen-
berg Foundation. J.C. receives support from the Re-
search Corporation for Science Advancement (Award
No. 21026) and the South Carolina Space Grant Con-
sortium. This work is based on observations made with
the Spitzer Space Telescope, which is operated by the
Jet Propulsion Laboratory, California Institute of Tech-
nology under a contract with NASA. This study made
use of the CDS services SIMBAD and VizieR, as well
as the SAO/NASA ADS service.
REFERENCES
Allard, F., Hauschildt, P. H., Alexander, D. R.,
Tamanai, A., & Schweitzer, A. 2001, ApJ, 556,
357, 357
Amara, A., & Quanz, S. P. 2012, MNRAS, 427, 948,
948
Baglin, A., Auvergne, M., Barge, P., et al. 2009, in
IAU Symposium, Vol. 253, IAU Symposium, ed.
F. Pont, D. Sasselov, & M. J. Holman, 71 -- 81
Bailey, V., Meshkat, T., Reiter, M., et al. 2014, ApJ,
780, L4, L4
Baraffe, I., Chabrier, G., Barman, T. S., Allard, F., &
Hauschildt, P. H. 2003, A&A, 402, 701, 701
Barnes, S. A. 2007, ApJ, 669, 1167, 1167
Barrado y Navascues, D. 1998, A&A, 339, 831, 831
Brandt, T. D., Kuzuhara, M., McElwain, M. W., et al.
2014, ApJ, 786, 1, 1
Bubenicek, J., Palous, J., & Piskunov, A. E. 1985, So-
viet Ast., 29, 625, 625
Burrows, A., Sudarsky, D., & Lunine, J. I. 2003, ApJ,
596, 587, 587
Butler, R. P., Marcy, G. W., Williams, E., et al. 1996,
PASP, 108, 500, 500
Carson, J. C., Eikenberry, S. S., Smith, J. J., & Cordes,
J. M. 2006, AJ, 132, 1146, 1146
Carson, J. C., Marengo, M., Patten, B. M., et al. 2011,
ApJ, 743, 141, 141
Chabrier, G., Baraffe, I., Allard, F., & Hauschildt, P.
2000, ApJ, 542, 464, 464
Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio,
F. A. 2008, ApJ, 686, 580, 580
Chauvin, G., Lagrange, A.-M., Bonavita, M., et al.
2010, A&A, 509, A52, A52
Chauvin, G., Vigan, A., Bonnefoy, M., et al. 2015,
A&A, 573, A127, A127
Cumming, A., Butler, R. P., Marcy, G. W., et al. 2008,
PASP, 120, 531, 531
Cutri, R. M., Skrutskie, M. F., van Dyk, S., et al. 2003,
VizieR Online Data Catalog, 2246, 0, 0
Dale, D. A., Bendo, G. J., Engelbracht, C. W., et al.
2005, ApJ, 633, 857, 857
Bergfors, C., Brandner, W., Daemgen, S., et al. 2013,
MNRAS, 428, 182, 182
Delorme, P., Lagrange, A. M., Chauvin, G., et al. 2012,
A&A, 539, A72, A72
Biller, B. A., Close, L. M., Masciadri, E., et al. 2007,
ApJS, 173, 143, 143
Dodson-Robinson, S. E., Veras, D., Ford, E. B., & Be-
ichman, C. A. 2009, ApJ, 707, 79, 79
Biller, B. A., Liu, M. C., Wahhaj, Z., et al. 2013, ApJ,
777, 160, 160
Bonavita, M., & Desidera, S. 2007, A&A, 468, 721,
721
Boss, A. P. 2006, ApJ, 637, L137, L137
Brandeker, A., Jayawardhana, R., Khavari, P., Haisch,
Jr., K. E., & Mardones, D. 2006, ApJ, 652, 1572,
1572
Dupuy, T. J., Liu, M. C., & Ireland, M. J. 2009, ApJ,
692, 729, 729
Duquennoy, A., & Mayor, M. 1991, A&A, 248, 485,
485
Ehrenreich, D., Lagrange, A.-M., Montagnier, G.,
et al. 2010, A&A, 523, A73, A73
Eisenbeiss, T., Ammler-von Eiff, M., Roell, T., et al.
2013, A&A, 556, A53, A53
15
Fazio, G. G., Hora, J. L., Allen, L. E., et al. 2004,
Kraus, A. L., Shkolnik, E. L., Allers, K. N., & Liu,
ApJS, 154, 10, 10
M. C. 2014b, AJ, 147, 146, 146
Fekel, F. C., Bopp, B. W., Africano, J. L., et al. 1986,
Lafreni`ere, D., Jayawardhana, R., & van Kerkwijk,
AJ, 92, 1150, 1150
M. H. 2010, ApJ, 719, 497, 497
Fortney, J. J., Marley, M. S., Saumon, D., & Lodders,
K. 2008, ApJ, 683, 1104, 1104
Lafreni`ere, D., Marois, C., Doyon, R., Nadeau, D., &
Artigau, ´E. 2007a, ApJ, 660, 770, 770
Fritz, T., Gillessen, S., Trippe, S., et al. 2010, MN-
Lafreni`ere, D., Doyon, R., Marois, C., et al. 2007b,
RAS, 401, 1177, 1177
ApJ, 670, 1367, 1367
Fuhrmann, K. 2004, Astronomische Nachrichten, 325,
Leconte, J., Soummer, R., Hinkley, S., et al. 2010, ApJ,
3, 3
-- . 2008, MNRAS, 384, 173, 173
Gaidos, E. J., Henry, G. W., & Henry, S. M. 2000, AJ,
120, 1006, 1006
Gliese, W., & Jahreiss, H. 1991, NASA STI/Recon
Technical Report A, 92, 33932, 33932
Holman, M. J., & Wiegert, P. A. 1999, AJ, 117, 621,
621
Ida, S., & Lin, D. N. C. 2004, ApJ, 604, 388, 388
Isella, A., Carpenter, J. M., & Sargent, A. I. 2009, ApJ,
701, 260, 260
Janson, M., Bonavita, M., Klahr, H., et al. 2011, ApJ,
736, 89, 89
Janson, M., Carson, J. C., Lafreni`ere, D., et al. 2012,
ApJ, 747, 116, 116
Janson, M., Quanz, S. P., Carson, J. C., et al. 2015,
A&A, 574, A120, A120
Jeffries, R. D., James, D. J., & Bromage, G. E. 1994,
MNRAS, 271, 476, 476
Juri´c, M., & Tremaine, S. 2008, ApJ, 686, 603, 603
Kasper, M., Apai, D., Janson, M., & Brandner, W.
2007, A&A, 472, 321, 321
King, J. R., Villarreal, A. R., Soderblom, D. R., Gul-
liver, A. F., & Adelman, S. J. 2003, AJ, 125, 1980,
1980
Koch, D. G., Borucki, W. J., Basri, G., et al. 2010, ApJ,
713, L79, L79
716, 1551, 1551
L´opez-Santiago, J., Montes, D., Crespo-Chac´on, I., &
Fern´andez-Figueroa, M. J. 2006, ApJ, 643, 1160,
1160
Makarov, V. V., Zacharias, N., & Hennessy, G. S.
2008, ApJ, 687, 566, 566
Makarov, V. V., Zacharias, N., Hennessy, G. S., Harris,
H. C., & Monet, A. K. B. 2007, ApJ, 668, L155,
L155
Maldonado, J., Mart´ınez-Arn´aiz, R. M., Eiroa, C.,
Montes, D., & Montesinos, B. 2010, A&A, 521,
A12, A12
Malo, L., Doyon, R., Lafreni`ere, D., et al. 2013, ApJ,
762, 88, 88
Mamajek, E. E., & Hillenbrand, L. A. 2008, ApJ, 687,
1264, 1264
Marengo, M., Megeath, S. T., Fazio, G. G., et al. 2006,
ApJ, 647, 1437, 1437
Marengo, M., Stapelfeldt, K., Werner, M. W., et al.
2009, ApJ, 700, 1647, 1647
Marley, M. S., Fortney, J. J., Hubickyj, O., Boden-
heimer, P., & Lissauer, J. J. 2007, ApJ, 655, 541,
541
Masciadri, E., Mundt, R., Henning, T., Alvarez, C.,
& Barrado y Navascu´es, D. 2005, ApJ, 625, 1004,
1004
Mason, B. D., Wycoff, G. L., Hartkopf, W. I., Dou-
glass, G. G., & Worley, C. E. 2001, AJ, 122, 3466,
3466
Kraus, A. L., Ireland, M. J., Cieza, L. A., et al. 2014a,
Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Mes-
ApJ, 781, 20, 20
senger, 114, 20, 20
16
Montes, D., L´opez-Santiago, J., Fern´andez-Figueroa,
M. J., & G´alvez, M. C. 2001a, A&A, 379, 976, 976
Tabernero, H. M., Montes, D., & Gonz´alez Hern´andez,
J. I. 2012, A&A, 547, A13, A13
Montes, D., L´opez-Santiago, J., G´alvez, M. C., et al.
Tetzlaff, N., Neuhauser, R., & Hohle, M. M. 2011,
2001b, MNRAS, 328, 45, 45
MNRAS, 410, 190, 190
Nakajima, T., & Morino, J.-I. 2012, AJ, 143, 2, 2
Naud, M.-E., Artigau, ´E., Malo, L., et al. 2014, ApJ,
787, 5, 5
Nielsen, E. L., & Close, L. M. 2010, ApJ, 717, 878,
878
Nielsen, E. L., Close, L. M., Biller, B. A., Masciadri,
E., & Lenzen, R. 2008, ApJ, 674, 466, 466
Thalmann, C., Desidera, S., Bonavita, M., et al. 2014,
A&A, 572, A91, A91
Torres, C. A. O., Quast, G. R., Melo, C. H. F., &
Sterzik, M. F. 2008, Young Nearby Loose Associa-
tions, ed. B. Reipurth, 757
van Leeuwen, F. 2007, A&A, 474, 653, 653
Veras, D., & Armitage, P. J. 2004, MNRAS, 347, 613,
Parker, R. J., & Quanz, S. P. 2012, MNRAS, 419,
613
2448, 2448
Veras, D., Crepp, J. R., & Ford, E. B. 2009, ApJ, 696,
Perets, H. B., & Kouwenhoven, M. B. N. 2012, ApJ,
1600, 1600
750, 83, 83
Perryman, M. A. C., Brown, A. G. A., Lebreton, Y.,
et al. 1998, A&A, 331, 81, 81
Plavchan, P., Werner, M. W., Chen, C. H., et al. 2009,
ApJ, 698, 1068, 1068
Pollacco, D. L., Skillen, I., Collier Cameron, A., et al.
2006, PASP, 118, 1407, 1407
Rafikov, R. R. 2007, ApJ, 662, 642, 642
Vican, L. 2012, AJ, 143, 135, 135
Vigan, A., Patience, J., Marois, C., et al. 2012, A&A,
544, A9, A9
Vogt, S. S., Marcy, G. W., Butler, R. P., & Apps, K.
2000, ApJ, 536, 902, 902
Werner, M. W., Roellig, T. L., Low, F. J., et al. 2004,
ApJS, 154, 1, 1
Wollert, M., Brandner, W., Reffert, S., et al. 2014,
Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954,
A&A, 564, A10, A10
954
Zuckerman, B., Bessell, M. S., Song, I., & Kim, S.
Riaz, B., Gizis, J. E., & Harvin, J. 2006, AJ, 132, 866,
2006, ApJ, 649, L115, L115
866
Zuckerman, B., Rhee, J. H., Song, I., & Bessell, M. S.
Riedel, A. R., Finch, C. T., Henry, T. J., et al. 2014,
2011, ApJ, 732, 61, 61
AJ, 147, 85, 85
Sandler, D. G., Stahl, S., Angel, J. R. P., Lloyd-Hart,
M., & McCarthy, D. 1994, Journal of the Optical
Society of America A, 11, 925, 925
Schlieder, J. E., L´epine, S., & Simon, M. 2012, AJ,
143, 80, 80
Soummer, R., Pueyo, L., & Larkin, J. 2012, ApJ, 755,
L28, L28
Spiegel, D. S., & Burrows, A. 2012, ApJ, 745, 174,
174
Zuckerman, B., & Song, I. 2004, ARA&A, 42, 685,
685
Zuckerman, B., Song, I., & Bessell, M. S. 2004, ApJ,
613, L65, L65
Zuckerman, B., Song, I., Bessell, M. S., & Webb, R. A.
2001, ApJ, 562, L87, L87
Zuckerman, B., Vican, L., Song, I., & Schneider, A.
2013, ApJ, 778, 5, 5
Stauffer, J. R., Hartmann, L. W., Prosser, C. F., et al.
1997, ApJ, 479, 776, 776
This 2-column preprint was prepared with the AAS LATEX macros
v5.2.
17
TABLE 3
TARGET SAMPLE PROPERTIES
HD
48189
113449
102647
174429
196982
197481
181296
216803
13507
217107
206860
35296
41700
173880
1237
17051
40979
75732
120136
179949
1835
222143
108799
30495
128987
19994
27045
197157
166
17925
197890
10008
77407
171488
130322
217014
92945
HIP
Other
RA
DEC
H Mag
Spectral Type
Distance (pc)
YMG
Age (Myr)
YMG / Age Reference
Binarity
31711
106231
63742
57632
23200
92680
102141
102409
95261
108706
113283
114252
10321
113421
107350
25278
28764
92161
1292
12653
28767
43587
67275
94645
1803
116613
60994
22263
71743
14954
117410
19990
102333
544
13402
37766
102626
7576
44458
91043
72339
113357
52462
LO Peg
GJ 182
GJ 4247
GJ 890
GJ 9839
GJ 285
EY Dra
V383 Lac
06 38 00.366
21 31 01.713
13 03 49.655
11 49 03.578
04 59 34.831
18 53 05.875
20 41 51.159
20 45 09.531
19 22 51.206
22 01 13.125
22 56 24.053
23 08 19.550
02 12 55.005
22 58 15.541
21 44 31.329
05 24 25.464
06 04 28.440
18 47 01.274
00 16 12.678
02 42 33.466
06 04 29.942
08 52 35.811
13 47 15.743
19 15 33.230
00 22 51.788
23 37 58.488
12 30 04.774
04 47 36.291
14 40 31.106
03 12 46.437
23 48 25.691
04 17 15.662
20 44 02.334
00 06 36.785
02 52 32.128
07 44 40.174
18 16 16.776
20 47 45.007
01 37 35.466
09 03 27.083
18 34 20.103
22 20 07.0258
14 47 32.727
22 57 27.980
10 43 28.272
-61 32 00.19
+23 20 07.37
-05 09 42.52
+14 34 19.41
+01 47 00.68
-50 10 49.88
-32 26 06.83
-31 20 27.24
-54 25 26.15
+28 18 24.87
-31 33 56.04
-15 24 35.80
+40 40 06.02
-02 23 43.38
+14 46 18.98
+17 23 00.72
-45 02 11.77
+18 10 53.47
-79 51 04.24
-50 48 01.06
+44 15 37.59
+28 19 50.95
+17 27 24.86
-24 10 45.67
-12 12 33.97
+46 11 57.96
-13 23 35.46
-16 56 04.04
-16 12 33.44
-01 11 45.96
-12 59 14.86
+20 34 42.93
-51 55 15.50
+29 01 17.40
-12 46 10.97
+03 33 08.84
+54 10 21.62
-36 35 40.79
-06 45 37.53
+37 50 27.53
+18 41 24.23
+49 30 11.763
-00 16 53.32
+20 46 07.79
-29 03 51.43
4.747
6.524
5.674
1.925
6.450
6.486
5.201
4.831
5.148
7.035
3.804
7.301
5.649
4.765
4.598
4.029
5.149
4.447
4.990
4.323
5.509
4.265
3.546
5.101
5.035
5.123
4.932
4.116
5.629
3.768
6.485
4.577
3.692
4.629
4.230
6.005
7.960
6.930
5.899
5.534
5.896
6.577
6.315
4.234
5.770
G1/G2V
K8
G5V
A3Vvar
M0Ve
K0Vp
M4Ve
M1Ve
A0Vn
M4V
K4V
M0Ve
G5V
G8IV
G0V
F8V
G0IV-V
A5III
G6V
G3IV
F8
G8V
F7V
F8V
G3V
G3/4V
G1/G2V
G3V
G6V
F8V
K5Vke
A3m
A9IV
K0V
K1V
M4.5Ve
dM1.5e
K0V
G5V
G0V
G0V
K1V
K0V
G5V
K1V
21.3
24.8
21.7
11.0
25.9
51.5
10.7
9.9
48.2
8.9
7.6
22.3
26.9
19.9
17.9
14.4
26.6
28.9
17.5
17.2
33.1
12.3
15.6
27.6
20.9
23.3
24.7
13.3
23.7
22.6
28.2
28.9
24.2
13.7
10.4
6.00
30.0a
52.2
24.0
30.5
38.0
27.5b
31.7
15.6
21.4
AB Dor
AB Dor
AB Dor
Argus
β Pictoris
β Pictoris
β Pictoris
β Pictoris
β Pictoris
Castor
Castor
Castor
Castor
Her-Lyr
Hyades
Hyades
Hyades
Hyades
Hyades
IC2391
IC2391
IC2391
Carina
Octans-Near
Octans-Near
Her-Lyr
LA
LA
LA
Tuc-Hor
LA
LA
LA
LA
LA
70 - 120
70 - 120
70 - 120
30 - 50
12 - 22
12 - 22
12 - 22
12 - 22
12 - 22
100 - 300
100 - 300
100 - 300
100 - 300
1000 - 10000
211 - 303
84 - 316
575 - 675
575 - 675
575 - 675
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
575 - 675
575 - 675
45 - 55
45 - 55
45 - 55
1000 - 10000
20 - 40
30 - 100
30 - 100
211 - 303
20 - 150
20 - 150
20 - 150
20 - 40
20 - 150
20 - 150
20 - 150
20 - 150
1000 - 10000
1000 - 10000
20 - 150
Z04a, M13
Z04a, L06, M13
Z04a, N12, M13
Y/0.6′′/810′′
Z11, M13
T8, S12
Z04b, T8
Z04b, T8
Z04b, T8
Z04b, M13
M01b, N12
M01b, N12
M01b, N12
M01b, N12
Y/2.2′′
Y/4600′′
Y/4.2′′
L06, F08
B07, M08, Mk08
Y/43.2′′
Y/705.2′′
M01b
B85
M01b
Y/192.4′′
Y/85.1′′
Y/1.8′′
Y/2.1′′
Y/2.1′′
Y/612′′
M01b, M10, N12
M01b, M10
N12
M10
M10
Z06
Z13
Z13
F04, L06
M01a, M10
M01b
J94
K14
M01b, M10
M01a, M01b
M01b
M01a, M01b
M01b, L06
1
8
YMG
LA
β Pictoris
Columba
Her-Lyr
TW Hydrae
Ursa Major
Ursa Major
Ursa Major
Ursa Major
Carina/Columba
β Pictoris
HD
HIP
Other
RA
DEC
H Mag
Spectral Type
Distance (pc)
TABLE 3 -- Continued
129333
181327
82443
116956
177724
29697
7590
217813
128311
147513
150706
175742
76644
82558
92139
112429
115383
141795
124498
220182
20630
130948
142
4208
10697
3651
20367
222404
27442
33636
39091
50554
75289
82943
92788
95128
114386
114783
192263
117176
134987
141937
143761
145675
71631
95270
46843
65515
93747
21818
5944
113829
71395
80337
80902
92919
44127
46816
51986
63076
64792
77622
69562
115331
15457
72567
522
3479
8159
3093
15323
116727
19921
24205
26394
33212
43177
47007
52409
53721
64295
64457
99711
65721
74500
77740
78459
79248
GJ 3789
14 39 00.210
19 22 58.943
09 32 43.759
13 25 45.533
19 05 24.608
04 41 18.856
01 16 29.253
23 03 04.977
14 36 00.560
16 24 01.289
16 31 17.585
18 55 53.225
08 59 12.454
09 32 25.568
10 37 18.140
12 55 28.548
13 16 46.516
13 31 46.617
15 50 48.966
14 14 21.357
23 21 36.513
03 19 21.696
14 50 15.811
00 06 19.175
00 44 26.651
01 44 55.825
00 39 21.806
03 17 40.045
23 39 20.852
04 16 29.029
05 11 46.448
05 37 09.892
06 54 42.825
08 47 40.390
09 34 50.737
10 42 48.528
10 59 27.973
13 10 39.824
13 12 43.786
20 13 59.846
13 28 25.809
15 13 28.667
15 52 17.547
16 01 02.662
16 10 24.314
+64 17 29.95
-54 32 16.97
+26 59 18.70
+56 58 13.78
+13 51 48.52
+20 54 05.45
+42 56 21.90
+20 55 06.87
+09 44 47.46
-39 11 34.71
+79 47 23.20
+23 33 23.93
+48 02 30.57
-11 11 04.70
-48 13 32.23
+65 26 18.51
+09 25 26.96
+29 16 36.72
+04 28 39.83
-15 21 21.76
+44 05 52.38
+03 22 12.72
+23 54 42.64
-49 04 30.68
-26 30 56.45
+20 04 59.34
+21 15 01.71
+31 07 37.36
+77 37 56.19
-59 18 07.76
+04 24 12.73
-80 28 08.84
+24 14 44.02
-41 44 12.45
-12 07 46.37
-02 11 01.52
+40 25 48.92
-35 03 17.21
-02 15 54.13
-00 52 00.77
+13 46 43.64
-25 18 33.65
-18 26 09.84
+33 18 12.63
+43 49 03.53
6.012
5.980
5.242
5.481
3.048
5.310
5.258
5.232
5.303
4.025
5.639
5.762
2.763
5.596
3.170
4.604
4.107
7.002
3.440
6.781
5.574
3.039
4.688
4.646
6.243
4.678
4.064
5.117
1.190
1.814
5.633
4.424
5.516
5.187
5.245
5.798
3.736
6.497
5.623
5.685
3.457
5.121
5.866
3.989
4.803
G1.5V
F5/F6V
G9V
G9V
A0Vn
K3V
G0
G5V
K0
G3/G5V
G0
K0V
A7Vn
K0V
F4IV
A5n
G0V
M4V
A2m
K4V
K1V
G5Vvar
G2V
F7V
G5V
G5IV
K0V
F8V
K1III
K2III
G0
G0V
F8V
F9VFe+0.3
F9VFe+0.5
G6V
G0V
K3V
K1V
K2.5V
G5V
G5V
G2/G3V
G2V
K0V
34.1
51.8
17.8
21.6
25.5
13.2
23.2
24.7
16.5
12.8
28.2
21.4
14.5
18.6
26.8
29.3
17.6
7.9c
21.6
30.2
21.5
9.14
18.2
25.7
32.4
32.6
11.0
26.7
14.1
18.2
28.4
18.3
29.9
29.2
27.5
35.5
14.0
28.9
20.5
19.3
18.0
26.2
32.3
17.2
17.6
Age (Myr)
YMG / Age Reference
M01a, M01b
Z01, Z04b
B14
F04, M10
N12
M01a, M01b
F04, M10
M01b, K03
K03
V12
F86, T11
P09
P09
M10, V12
R14
V12
C10, M13
G00, B07, M08
M08 , M10, V12
D09
20 - 150
12 - 22
20 - 40
211 - 303
8 - 12
400 - 600
400 - 600
400 - 600
1000 - 10000
1000 - 10000
1000 - 10000
400 - 600
450 - 1050
50 - 75
50 - 150
50 - 450
130 - 160
20 - 40
220 - 820
12 - 10000
200-318
350-700
640-1001
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
Binarity
Y/0.8′′
Y/64.7′′
Y/1494.63′′
Y/345′′
Y/0.7′′/2.4′′
Y/0.5′′
Y/0.1′′/2.6′′
Y/4.1′′
Y/42.9′′
Y/0.9′′
Y/13.1′′
Y/21.5′′
1
9
TABLE 3 -- Continued
HD
HIP
Other
RA
DEC
H Mag
Spectral Type
Distance (pc)
YMG
Age (Myr)
YMG / Age Reference
Binarity
160691
216435
210277
216437
74575
59967
73350
37124
52265
36705
21845
102077
105963
139084
172555
202730
218738
51849
141272
155555
220140
11131
43162
160934
9826
13445
46375
162020
186427
190360
86796
113044
109378
113137
42828
36515
42333
26381
33719
25647
16563
57269
59432
76629
92024
105319
114379
33560
77408
84586
115147
8486
29568
86346
7513
10138
31246
87330
96901
98767
AF Hor
17 44 08.701
22 53 37.932
22 09 29.866
22 54 39.482
08 43 35.538
07 30 42.512
08 37 50.294
05 37 02.486
07 00 18.036
02 41 47.31
05 28 44.830
03 33 13.491
11 44 38.463
12 11 27.754
15 38 57.543
18 45 26.900
21 19 51.990
23 09 57.372
06 58 26.051
15 48 09.463
17 17 25.505
23 19 26.633
01 49 23.356
06 13 45.296
17 38 39.634
01 36 47.842
02 10 25.934
06 33 12.622
17 50 38.355
19 41 51.972
20 03 37.406
-51 50 02.59
-48 35 53.83
-07 32 55.15
-70 04 25.35
-33 11 10.99
-37 20 21.70
-06 48 24.78
+20 43 50.84
-05 22 01.78
-52 59 30.7
-65 26 54.86
+46 15 26.53
-49 25 02.75
+53 25 17.45
-57 42 27.34
-64 52 16.54
-53 26 57.93
+47 57 30.13
-12 59 30.58
+01 34 18.27
-66 57 03.73
+79 00 12.67
-10 42 12.86
-23 51 42.98
+61 14 16.03
+41 24 19.64
-50 49 25.42
+05 27 46.53
-40 19 06.07
+50 31 03.08
+29 53 48.49
3.724
4.687
4.957
4.923
4.227
5.253
5.318
6.021
5.033
7.851
4.845
6.457
6.642
5.867
5.994
4.251
4.224
5.788
6.368
5.610
4.907
5.512
5.289
4.863
6.998
2.957
4.245
6.072
6.649
4.695
4.239
G5V
G0V
G8V
G1VFe+0.3
B1.5III
G3V
G5V
G4IV-V
G0V
M2Ve
K0V
K2
K1V
K0V
K0V
A7V
A5V
G5Ve
K4V
G8V
G5IV
G9V
G1Vk
G5V
K7Ve
F8V
K0V
K1IV
K3V
G5V
G7IV-V
15.5
32.6
21.6
26.8
269.5
21.8
24.0
33.7
29.0
27.0d
15.2
34.4
48.6
30.2
38.5
28.6
30.3
23.7
21.7
21.3
31.5
19.2
22.6
16.7
33.1
13.5
10.8
34.8
29.4
21.2
15.9
Castor
Hyades
Tuc-Hor
AB Dor
AB Dor
β Pictoris
β Pictoris
LA
β Pictoris
Ursa Major
IC 2391
AB Dor
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
13.2 - 18.4
100 - 300
575 - 675
1000 - 10000
1000 - 10000
20 - 40
70 - 120
70 - 120
30 - 120
242 - 302
12 - 22
12 - 22
50 - 900
12.7 - 48
30 - 220
20 - 150
12 - 22
16 - 50
400 - 600
45 - 55
70 - 120
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
1000 - 10000
T11
N12
M10, T12
M13, Z04b
Z04a, L06, M13
Z04a, L06, M13
W14
B07, M08
Z04b, M13
Z04b, M13
P09, V12
P09, T11
P09, T11
M01b, M10
Z04b, M13
Mk07, T11
M01b, M10, N12
M10, N12
Z04a, L06, M13
Y/22.1′′
Y/8.9′′
Y/9.6′′
Y/0.2′′
Y/13.5′′
Y/10.2′′
Y/71.4′′
Y/7.3′′
Y/15.8′′
Y/0.7′′
Y/17.9′′
Y/34.04′′
Y/10.8′′/962.6′′
Y/192.9′′
Y/24.6′′/164′′
Y/0.12′′
Y/55.4′′
Y/2.4′′
Y/11.2′′
Y/41.6′′
Y/178.2′′
2
0
NOTE. -- Target right ascension (RA) and declination (DEC) values are taken from the revised Hipparcos catalogue (van Leeuwen 2007) at equinox=J2000, epoch=J2000 computed by Vizier.
H band magnitudes are taken from 2MASS All-Sky Catalog of Point Sources (Cutri et al. 2003). Distances taken from revised Hipparcos catalogue unless noted otherwise with indices [a,b,c,d], relevant
references given below. Column 9 notes YMG membership. Age estimates for YMG targets are taken from estimates of the group age given in table 2. Targets with unconstrained ages are assigned an age
of 1 - 10 Gyr. Column 11 notes literature sources identifying YMG membership, or the sources estimating age for several young targets with no known YMG membership. Column 12 notes Y for targets
identified as part of a binary system and gives the separation of the binary companion, identification and separation values are taken from the Washington double star catalogue (Mason et al. 2001).
a(Plavchan et al. 2009)
b(Montes et al. 2001a)
c(Gliese & Jahreiss 1991)
d(Riaz et al. 2006)
References. -- (B07) (Barnes 2007); (B14) (Brandt et al. 2014); (B85) (Bubenicek et al. 1985); (C10) (Chauvin et al. 2010); (D09) (Dupuy et al. 2009); (F04, F08) (Fuhrmann 2004, 2008); (F86)
(Fekel et al. 1986); (G00) (Gaidos et al. 2000); (J94) (Jeffries et al. 1994); (K03) (King et al. 2003); (K14) (Kraus et al. 2014b); (L06) (L´opez-Santiago et al. 2006); (M01a, M01b) (Montes et al. 2001a,b);
(M08) (Mamajek & Hillenbrand 2008); (M10) (Maldonado et al. 2010); (M13) (Malo et al. 2013); (Mk07, Mk08) (Makarov et al. 2007, 2008); (N12) (Nakajima & Morino 2012); (P09) (Plavchan et al.
2009) ; (R14) (Riedel et al. 2014); (S12) (Schlieder et al. 2012); (T8) (Torres et al. 2008); (T11) (Tetzlaff et al. 2011); (T12) (Tabernero et al. 2012); (V12) (Vican 2012); (W14) (Wollert et al. 2014); (Z01)
(Zuckerman et al. 2001); (Z04a, Z04b) (Zuckerman et al. 2004; Zuckerman & Song 2004); (Z06, Z11, Z13) (Zuckerman et al. 2006, 2011, 2013).
2
1
|
0912.3356 | 1 | 0912 | 2009-12-17T10:18:25 | Dust Properties of Protoplanetary Disks in the Taurus-Auriga Star Forming Region from Millimeter Wavelengths | [
"astro-ph.EP"
] | We present the most sensitive 3 mm-survey to date of protoplanetary disks carried in the Taurus-Auriga star forming region (average rms of about 0.3 mJy), using the IRAM PdBI. With our high detection rate of 17/19, we provide the first detections at wavelengths longer than about 1 mm for 12 sources. This enables us to study statistically the mm SED slopes and dust properties of faint disks and compare them to brighter disks using a uniform analysis method. With these new data and literature measurements at sub-millimeter and millimeter wavelengths, we analyze the dust properties of a sample of 21 isolated disks around T Tauri stars in the Taurus-Auriga star forming region. Together with the information about the disks spatial extension from sub/mm-mm interferometric studies, we derive from the observed sub-mm/mm spectral energy distribution constraints on the dust opacity law at these wavelengths, using two-layer flared disk models and a self-consistent dust model that takes properly into account the variation of the dust opacity with grain growth. We find evidence for the presence in the disk midplane of dust particles that have grown to sizes as large as at least 1 millimeter in all the disks of our sample, confirming what was previously observed on smaller brighter objects. This indicates that the dust coagulation from ISM dust to mm-sized grains is a very fast process in protoplanetary disks, that appears to occur before a young stellar object enters the Class II evolutionary stage. Also, the amount of these large grains in the disk outer regions is stationary throughout all the Class II evolutionary stage, indicating that mechanisms slowing down the dust inward migration are playing an important role in the Taurus-Auriga protoplanetary disks. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. 6439
May 20, 2018
c(cid:13) ESO 2018
9
0
0
2
c
e
D
7
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
5
3
3
.
2
1
9
0
:
v
i
X
r
a
Dust Properties of Protoplanetary Disks in the Taurus-Auriga
Star Forming Region from Millimeter Wavelengths
L. Ricci1, L. Testi1,2, A. Natta2, R. Neri3, S. Cabrit4, and G. J. Herczeg5
1 European Southern Observatory, Karl-Schwarzschild-Strasse 2, D-85748 Garching, Germany
2 INAF - Osservatorio Astrofisico di Arcetri, Largo Fermi 5, I-50125 Firenze, Italy
3 Institut de Radioastronomie Millim´etrique, 300 Rue de la Piscine, F-38406 Saint Martin d'H`eres, France
4 LERMA, UMR 8112 du CNRS, Observatoire de Paris, 61 Av. de l'Observatoire, F-75014 Paris, France
5 Max Planck Institut fur Extraterrestrische Physik, Postfach 1312, D-85741 Garching, Germany
Received 2009 October 5/ Accepted 2009 December 15
ABSTRACT
We present the most sensitive 3 mm-survey to date of protoplanetary disks carried in the Taurus-Auriga star forming
region (average rms of about 0.3 mJy), using the IRAM PdBI. With our high detection rate of 17/19, we provide the
first detections at wavelengths longer than about 1 mm for 12 sources. This enables us to study statistically the mm
SED slopes and dust properties of faint disks and compare them to brighter disks using a uniform analysis method.
With these new data and literature measurements at sub-millimeter and millimeter wavelengths, we analyze the dust
properties of a sample of 21 isolated disks around T Tauri stars in the Taurus-Auriga star forming region. Together
with the information about the disks spatial extension from sub/mm-mm interferometric studies, we derive from the
observed sub-mm/mm spectral energy distribution constraints on the dust opacity law at these wavelengths, using
two-layer flared disk models and a self-consistent dust model that takes properly into account the variation of the dust
opacity with grain growth. We find evidence for the presence in the disk midplane of dust particles that have grown to
sizes as large as at least 1 millimeter in all the disks of our sample, confirming what was previously observed on smaller
brighter objects. This indicates that the dust coagulation from ISM dust to mm-sized grains is a very fast process in
protoplanetary disks, that appears to occur before a young stellar object enters the Class II evolutionary stage. Also,
the amount of these large grains in the disk outer regions is stationary throughout all the Class II evolutionary stage,
indicating that mechanisms slowing down the dust inward migration are playing an important role in the Taurus-Auriga
protoplanetary disks. Another result is that the spectral index between 1 and 3 mm for the 6 faintest disks in our sample
is on average smaller than for the brighter disks, indicating either that these fainter, yet unmapped, disks are spatially
much less extended than the brighter spatially resolved disks, or that fainter disks have typically larger dust grains in
their outer regions. Considering that these fainter disks are more representative of the bulk of the disk population than
the brighter ones, this may have important consequences for the theories of planetesimal formation and disk formation
and evolution. Finally, we investigate the relations between the derived dust properties, namely dust mass and grain
growth, and the properties of the central star, like its mass, age and mass accretion rate.
Key words. stars: planetary systems: protoplanetary disks -- stars: planetary systems: formation -- stars: formation
1. Introduction
Circumstellar disks play a fundamental role in the physical
processes involved in star and planet formation. In these
systems part of the material (gas and dust) loses angu-
lar momentum and accretes onto the central forming star,
while another part of it may give birth to a planetary sys-
tem. Although the content of dust is only a small fraction
of the overall material in a circumstellar disk, dust grains
are crucial elements in the early stages of planet forma-
tion. According to the core-accretion scenario (Safronov &
Zvjagina 1969, Pollack et al. 1996), growth from an ini-
tial ISM population of submicrometer-sized dust grains to
planetesimals of 1 − 100 km sizes is believed to be the key
mechanism of early planet formation. These planetesimals
"embryos" then continue to grow leading to the formation
of at least terrestrial planets and possibly the cores of giant
planets.
Observational evidence of grain growth from the ISM
dust has been obtained from a variety of techniques at dif-
ferent wavelengths. Throop et al. (2001) observed in optical
and near-infrared the translucent edge of the 114-426 pro-
toplanetary disk in the Orion Nebula Cluster, and derived
from the extinction curve of the background nebular emis-
sion a lower limit of a few microns for the dust grains in the
very outer regions of the disk, about 500 AU away from the
central star. Another method to probe the presence of large
grains in the disk is through the shape and intensity of the
10- and 20-µm silicate features. The data indicate a large
variety of silicate profiles, ranging from strongly peaked
silicate bands and steeply rising spectral energy distribu-
tions (SEDs) to "boxy" silicate profiles and flatter SEDs
(ISO, Malfait et al. 1998; Spitzer Space Telescope, Kessler-
Silacci et al. 2006). The boxy features with low feature-to-
continuum ratios are interpreted as grain growth to micron
size (Bouwman et al. 2001).
However, the techniques outlined so far can only probe
dust grains as large as some µm in the disk surface layers.
They are not sensitive to larger grains in the disk mid-
2
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
plane, the region in which planet formation is supposed to
occur. In order to probe larger grains in the midplane, ob-
servations at millimeter wavelengths are needed. Beckwith
& Sargent (1991) showed that T Tauri stars have shallow
SEDs at sub-millimeter wavelengths. Under the assumption
of optically thin emission at these frequencies, this implies a
dust opacity dependence on wavelength (κmm ∝ λ−β) much
flatter than in the ISM (βISM = 1.7), which is naturally in-
terpreted in terms of grain growth (see e.g. Draine 2006).
However, this interpretation of the disk SEDs is not unique,
since the same data can be explained by very small opti-
cally thick disks. To break this degeneracy and sort out
the effect of potentially large optical depth it is neces-
sary to spatially resolve the disks to determine their actual
sizes. Furthermore, in this context observations at millime-
ter wavelengths are very useful, since at these lower fre-
quencies the impact of optically thick disk inner regions
to the total emission is expected to be lower. Therefore,
to actually probe the dust properties in protoplanetary
disks, one needs to combine the determination of the sub-
millimeter/millimeter SED with information on the disk
extension from high-angular resolution interferometric ob-
servations (see e.g. Testi et al. 2001).
In the last years, several sub-mm/mm interferometric
observations have been carried out to investigate dust grain
growth in protoplanetary disks. Wilner et al. (2000) re-
solved the disk around the TW Hya pre-main-sequence
(PMS) star at 7 mm using the Very Large Array (VLA).
Extensive modelling of the SED of this source has shown
that the dust grains in the outer parts of the TW Hya
disk have grown to at least ∼ 1 cm (Calvet et al. 2002).
An analogous result has been obtained for the disk around
CQ Tau (Testi et al. 2003). More recently, Rodmann et
al. (2006) resolved 10 disks in the Taurus-Auriga star form-
ing region and found clear evidence of grain growth in 4 of
them. Lommen et al. (2007) resolved 1 disk in Chamaeleon
and 4 in Lupus with the Submillimeter Array (SMA) at
1.4 mm and with the Australia Telescope Compact Array
(ATCA) at 3.3 mm, and found clear evidence of dust grain
growth to sizes of a few millimeter for 4 of them. Finally
Schaefer et al. (2009) observed a sample of 23 low mass
PMS stars (with spectral types of K7 and later) in Taurus-
Auriga at 1.3 mm and 2.6 mm with the Plateau de Bure
Interferometer (PdBI); they detected only 8 sources at 1.3
mm and 6 at 2.6 mm, and found evidence of grain growth
for the 3 disks that they could spatially resolve.
In this paper we present our analysis on 21 protoplane-
tary disks around T Tauri stars in the Taurus-Auriga star
forming region without stellar companions in the range
0.05′′ −3.5′′. For 11 of these objects, mainly faint disks with
F1.3mm < 100 mJy, we have obtained new data at ∼ 3 mm
with the Plateau de Bure Interferometer1 (hereafter PdBI)
with an average rms of about 0.3 mJy. Together with the
data already present in the literature at sub/mm- and mm-
wavelengths, and with the information obtained in the last
years on the disks spatial extension from high-angular reso-
lution interferometric observations at mm wavelengths, we
investigate the dust properties in the disks, namely grain
growth and dust mass, and their relation with the proper-
ties of the central star.
1 The Plateau de Bure Interferometer at IRAM is supported
by INSU/CNRS (France), MPG (Germany) and IGN (Spain).
In Section 2 we present our new PdBI data, the proper-
ties of our sample, the method we used to estimate the main
stellar physical quantities, and the data we used for our
analysis. In Section 3 we describe the disk models adopted
for the analysis of the disks sub-mm/mm SED. In Section
4 we show and discuss the results of our study in terms of
dust grain growth and dust mass, whereas in Section 5 we
summarize our main findings.
2. New 3mm observations and sample properties
2.1. New PdBI observations
We observed a total list of 19 targets; 16 of them were se-
lected to be relatively faint, with 15 mJy< F1.3mm < 100
mJy and with no known detections beyond about 1mm;
the other 3 (HL, DO and DR Tau) are brighter sources
that were observed for references purposes. The observa-
tions were carried out with PdBI between July and August,
2007. Owing to antenna maintenance work, all observations
were carried out in a subarray of the six-element interfer-
ometer. The antennas were arranged in very compact con-
figurations that provided sensitive baselines from 15 m to
80 m. The dual-polarization receivers, which were observ-
ing in the 3mm band, had typical receiver temperatures of
about 35 K and were providing image sideband rejection
values better than 10 dB. To achieve maximum sensitivity
in the continuum, the spectral correlator was adjusted to
cover a total effective bandwidth of about 2 GHz.
The observing procedures were set up to fill in gaps
in the scheduling without any specific tuning wavelength
within the 3mm band. Visibilities were obtained using on-
source integration times of 22 min interspersed with 2 min
calibrations on a nearby calibrator. Each star was observed
on one or more occasions for a minimum of 40 min and a
maximum of 2 hr on-source. The atmospheric phase stabil-
ity on the baselines was always better than 40◦, consistent
with seeing conditions of 1′′-3′′, typical for summer. The ab-
solute flux density scale was calibrated on 3C84, 0528+134,
MWC349 and Mars, and was found to be accurate to 10%.
The receiver passband shape was determined with excellent
precision on a strong calibrator like 3C84.
We used the GILDAS package for the data reduction
and analysis. The data were calibrated in the antenna based
manner. The continuum visibilities were gridded with nat-
ural weighting and no taper to maximize the sensitivity.
Since none of the target stars was found to be resolved in
synthesized beams of 2′′-4′′, a point source model was fit-
ted to the calibrated visibilities of each star to estimate the
flux density of the continuum in the 3mm band. The results
are summarized in Table 1. Among the 19 targets, 17 were
detected, including 14 of the 16 faint sources, thanks to our
good sensitivity; 12 of these are first detections longward of
about 1mm, strongly expanding the available database of
3mm data towards fainter disks. The 2.6mm detections of
LkHa 358 and GO Tau were recently published by Schaefer
et al. (2009).
In Sections 2.2, 2.3 and 2.4 we describe the final sample
used for our analysis of dust properties, the adopted method
to estimate the stellar properties and mass accretion rates
summarized in Table 2, whereas in 2.5 we report the sub-
mm/mm data that we collected from the literature for our
analysis.
Table 1. Summary of the new PdBI observations.
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
3
Object
α (J2000)
δ (J2000)
ν
(GHz)
λ
(mm)
Fν
(mJy)
rms
(mJy)
Comments1
CW Tau
CX Tau
DE Tau
DK Tau
DO Tau
DP Tau
DR Tau
DS Tau
FM Tau
FP Tau
FY Tau
FZ Tau
GK Tau
GO Tau
Haro 6-28
HL Tau
HO Tau
LkHa 358
SU Aur
04:14:17.0
04:14:47.8
04:21:55.6
04:30:44.3
04:38:28.6
04:42:37.7
04:47:06.2
04:47:48.6
04:14:13.5
04:14:47.4
04:32:30.5
04:32:31.8
04:33:34.6
04:43:03.1
04:35:56.9
04:31:38.4
04:35:20.2
04:31:36.2
04:55:59.4
28:10:56.51
26:48:11.16
27:55:05.55
26:01:23.96
26:10:49.76
25:15:37.56
16:58:43.05
29:25:10.96
28:12:48.84
26:46:26.65
24:19:56.69
24:20:02.53
24:21:06.07
25:20:17.38
22:54:36.63
18:13:57.37
22:32:13.98
18:13:43.20
30:34:10.39
84.2
86.2
92.6
84.2
84.2
84.2
100.8
100.8
100.8
100.8
100.8
100.8
100.8
86.2
87.0
84.5
87.0
84.5
84.5
3.44
3.57
1.01
3.49
3.32
3.23
4.70
3.57
3.57
13.80
3.57 < 0.78
13.90
2.97
2.97
2.96
2.65
2.97
11.80
2.97
2.97
1.90
2.97
2.89
2.97 < 0.96
4.05
3.49
0.92
3.45
3.53
43.40
2.03
3.45
2.46
3.53
3.53
2.88
0.31
0.13
0.25
0.20
0.29
0.26
0.47
0.21
0.24
0.26
0.28
0.24
0.32
0.34
0.18
0.68
0.53
0.22
0.19
2.5′′ binary
no sub-mm info
no sub-mm info
no sub-mm info
2.5′′ binary
0.7′′ binary
flat spectrum
no sub-mm info
1) Reason why the source has not been considered in the analysis (see text for more details).
2.2. Sample
Our final sample of 21 sources consists of all young stellar
objects (YSO) in Taurus-Auriga catalogued in Andrews &
Williams (2005) that fulfill the following criteria: (1) a Class
II infrared SED, in order to avoid contamination of sub-mm
fluxes by a residual envelope; (2) information on the central
star through optical-NIR spectroscopic/photometric data,
necessary to calculate self-consistent disk SED models; (3)
one detection at ∼ 3mm (either from our new PdBI obser-
vations or from the literature) and at least one detection in
the 0.45 mm < λ < 0.85 mm spectral region. This selection
criterion was chosen in order to probe a broad enough spec-
tral window to efficiently constrain the optically thin part
of the sub-mm/mm spectral energy distribution (SED); (4)
no evidence of stellar companions in the 0.05−3.5′′ range in
angular separation. This range corresponds to about 5−500
AU in projected physical separation at the Taurus-Auriga
star forming region distance, here assumed to be 140 pc for
all the sources of our sample (Bertout et al. 1999). The limit
of 3.5 arcsec ensures that mm interferometric flux measure-
ments, of typical resolution 1 arcsec, are not contaminated
by the companion's disk, while binaries closer than 5 AU
should have outer disk properties similar to those of single
stars on the scales probed by (sub-)mm data.
Only
by Andrews &
Williams (2005) fulfill the above 4 criteria. They are
catalogued
21
sources
listed in Table 2, and include 11 sources from our PdBI
sample2.
Figure 1 reports histograms that describe the com-
pleteness level of our final sample with respect to all
Taurus-Auriga Class II YSOs catalogued in Andrews &
Williams (2005), and fulfilling the "isolation" criterion (4).
Our sample contains 63% of these "isolated" class II show-
ing a flux at ∼ 0.85mm greater than 100 mJy, and 31% of
the sources with lower 0.85mm fluxes. Therefore, our sam-
ple is not complete even for the brightest YSOs at 0.85mm.
This is due to a lack of observations at ∼ 3mm for these
sources. However, thanks to the high sensitivity of our new
PdBI observations (see Section 2.1) our sample enables us
to study statistically for the first time the dust properties
2 The 8 sources dropped from the PdBI sample are HL Tau
(a flat spectrum source that retains an envelope; Padgett et
al. 1999); DP Tau, FP Tau, FY Tau and LkHa 358 that have
no available submm detections for SED fitting; and DK Tau,
GK Tau and Haro 6-28 which are binary systems with an-
gular separations of 2.5′′ (Simon et al. 1992), 2.5′′ (Reipurth
& Zinnecker 1993) and 0.7′′ (Leinert et al. 1993) respectively.
We keep in our sample the wide binaries DS Tau and HO Tau
(angular separations of 7.1′′ and 6.9′′ respectively, Mathieu et
al. 1994); they do not show any emission from the compan-
ions from our new 3 mm interferometric observations. The other
dropped multiple systems from the AW05 catalog are DH Tau
(2.3′′ binary, Itoh et al. 2005) , GG Tau (quadruple system with
a circumbinary ring of inner and outer radii of 180 and 260 AU
respectively, Pinte et al. 2007), and UY Aur (0.9′′ binary, Leinert
et al. 1993)
4
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
of faint disks (i.e. with F1.3mm < 100 mJy, F0.85mm < 100
mJy) and compare them to the brighter ones using a uni-
form analysis method. In terms of the stellar properties, it is
important to note that the Andrews & Williams (2005) cat-
alogue includes only a few very low mass PMS stars: only
stars with an estimated mass higher than about 0.2 M⊙
have been detected because of the sensitivity limits of the
current facilities at sub-mm and mm wavelengths that make
the detection of disks around very low mass YSOs very diffi-
cult. The sensitivity limitation is even more severe at 3mm,
therefore our final sample includes 58% of all the class II
isolated sample of Andrews & Williams (2005) with an esti-
mated mass greater than 0.4 M⊙, but only two sources with
M⋆ < 0.4 M⊙ (out of 10 in Andrews & Williams 2005).
2.3. Stellar properties
Spectral types for all the sources in our sample are taken
from the literature (Kenyon & Hartmann 1995, Briceno et
al. 1998) and cover a range from G2 (SU Aur) to M2.5
(CX Tau). Spectral types of M0 and earlier have been con-
verted to effective temperatures with the dwarf tempera-
ture scale of Schmidt-Kaler (1982), whereas for types later
than M0 we adopted the temperature scale developed by
Luhman 1999. Typical uncertainties for these sources are
∼ 1 − 1.5 in spectral sub-types or ∼ 100 − 150 K in tem-
perature.
The stellar luminosities of all the pre-main-sequence
(PMS) stars were computed from their J-band flux (2MASS
All Sky Catalog of point sources, Cutri et al. 2003) to min-
imize contamination from UV and IR excess emission. For
the bolometric corrections we adopted the dwarf values
from Kenyon & Hartmann (1995) that are considered to
be satisfactory approximations for these young sources in
the J-band (Luhman 1999).
In order to estimate the amount of extinction toward
these young sources, we calculated the color excesses using
intrinsic colors provided by Kenyon & Hartmann (1995) for
G2−K7 spectral types and by Leggett (1992) for M0−M2.5
types. To ensure that the color excess reflects only the ef-
fect of reddening, minimizing the emission due to accretion,
the typical selected colors are between the R and H bands.
Since for some of the objects at later spectral types in our
sample R − I measurements are not available, we dered-
dened the J − H and H − Ks colors from the 2MASS All
Sky Catalog of point sources3 to the locus observed for clas-
sical T Tauri stars (CTTS) by Meyer et al. (1997), following
the method described in Briceno et al. (2002). Extinctions
are finally calculated adopting the extinction law of Rieke
& Lebofsky (1985).
The computed photospheric luminosities cover a range
from 0.16 L⊙ (HO Tau) to 10.6 L⊙ (SU Aur). Considering
the uncertainties in the photometry, reddenings, bolometric
corrections and distance, the typical errors in the bolomet-
ric luminosities are ±0.08 − 0.13 in log L⋆. Slight differ-
ences from the values obtained by other authors are typi-
cally within the uncertainties and are not significant for the
purposes of this paper.
3 For
this analysis
the JHKs magnitudes have been
transformed from the 2MASS photometric system to the
Johnson-Glass one using the color transformations reported in
Carpenter (2001) and Bessell & Brett (1988).
Fig. 2. H-R diagram for the sources of our sample. The
dashed lines correspond to the isochrones for ages in Myr
as labeled at the right end of the lines from the 1999 PMS
evolutionary models, while the solid ones represent the evo-
lutionary tracks from the same models for PMS stars with
masses as labeled at the top of the evolutionary track lines.
In this diagram the evolutionary tracks start from an age
of 0.1 Myr. The errorbars for the (Teff, L⋆) values of our
objects are not shown, the uncertainties in the photomet-
ric parameters are typically ∼ 0.08 − 0.13 in log L⋆ and
∼ 100 − 150 K in temperature.
Given the effective temperatures and photospheric lu-
minosities, we placed the PMS stars on the H-R diagram
and derived stellar masses and ages adopting the theoretical
tracks and isocrones of Palla & Stahler (1999) (see Figure
2). One important source of uncertainty for stellar masses
and ages is given by the spread of values obtainable by
using different PMS evolutionary models. For example, if
compared to the ones obtained with the Baraffe et al. (1998)
models, our adopted values of stellar masses and ages are
typically lower by a factor of ∼ 1.5 and ∼ 2 respectively.
Our choice of using the Palla & Stahler (1999) models is
uniquely due to the more complete coverage of the HR di-
agram plane, that allows us to use the same evolutionary
models for all the sources of our sample. One should al-
ways bear in mind the high uncertainty associated to these
quantities, expecially to the stellar age (see Section 4.1 for a
more deailed discussion about the estimates of YSO ages).
According to the Palla & Stahler (1999) models the ranges
spanned by our sample go from about 0.3 M⊙ (CX Tau)
to 2.2 M⊙ (RY Tau) in stellar masses and from about 0.1
Myr (UZ Tau E) to 17 Myr (DS Tau) in stellar ages.
The main stellar quantities described here are summa-
rized in Table 2.
2.4. Mass accretion rates
Given the stellar luminosities and effective temperatures
we calculated the stellar radii using the Stefan-Boltmann
law. Once the stellar mass M⋆ and radius R⋆ are set, the
mass accretion rate Macc can be directly obtained from an
estimate of the accretion luminosity Lacc by
Macc = 1.25
R⋆Lacc
GM⋆
,
(1)
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
5
Fig. 1. Histograms higlighting some properties of our selected sample. In all the histograms our sample is represented
by black+grey columns, the black columns represent the disks in our sample for which we obtained new PdBI data at ∼
3 mm, while the total columns (black+grey+white) include all the class II YSOs from the Andrews & Williams (2005)
catalogue with no evidence of stellar companions in the 0.05 − 3.5′′ interval in angular separation, from which our sample
has been selected (see text). Upper left histogram represents the distribution of the ∼0.85 mm fluxes, including upper
limits, for all the sources but two, 04113+2758 and GM Aur, that have not been observed at ∼0.85 mm. Upper right and
bottom histograms show the distribution of stellar spectral type and estimated stellar mass (see Section 2.3) respectively.
Four class II YSOs are not included in these histograms because for them an estimate of the stellar spectral type is not
available. These are 04301+2608, FT Tau, Haro 6-39 and IC 2087.
where G is the gravitational constant and the factor of ∼
1.25 is calculated assuming a disk truncation radius of ∼
5 R⋆ (Gullbring et al. 1998).
The estimates of Lacc were obtained from spectroscopic
detections of excess Balmer continuum emission in the lit-
erature (see references in Table 2). These estimates are
obtained primarily from measuring the excess flux in the
Balmer continuum shortward of 3646 A. The luminos-
ity is then obtained by applying a bolometric correction
calculated from shock models (Calvet & Gullbring 1998,
Gullbring et al. 2000) or simplistic plane-parallel slabs
(Valenti et al. 1993, Gullbring et al. 1998).
Two sources in our sample, FZ Tau and GO Tau,
do not have literature estimates of accretion luminosity
from the excess Balmer continuum emission. We obtained
the accretion luminosity for these two sources from low-
resolution spectra from 3200 -- 8700 A obtained with the
Double Spectrograph at the 5m Hale Telescope at Palomar.
The accretion luminosity was then calculated from the ex-
cess Balmer continuum emission, following the method de-
scribed in Herczeg & Hillenbrand (2008).
2.5. Disks sub-mm and mm data from the literature
In order to probe the sub-millimetre and millimeter SED
we collected data of the dust continuum emission between
∼ 0.450 to ∼ 7 mm from several works in the literature,
listed in Table 3. These data, except for 7 mm measure-
ments, have been used together with the new PdBI ob-
servations at ∼ 3 mm to constrain the disk properties by
fitting the sub-mm/mm SED with the models described
in Section 3. The 7-mm data is excluded because free-free
emission from ionized gas contaminates disk emission, with
free-free emission typically contributing ∼ 20% of the 7mm
flux (Rodmann et al. 2006). Simultaneous cm observations
are required to constrain the free-free emission at 7mm,
which is variable on short timescales (<
∼ a few days, see
Lommen et al. 2009). Only 2 disks in our sample (RY
Tau and UZ Tau E) have been observed nearly simultane-
ously at centimeter and millimeter wavelengths (Rodmann
et al. 2006), without which the 7mm fluxes should be con-
sidered only as upper limits for the dust emission. We there-
fore did not include the 7mm data in our analysis, although
we included the available 7mm data from the literature in
Figure 4 and verified a posteriori that the fluxes of our disk
6
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
Table 2. Stellar properties.
Object
ST
Teff
(K)
L⋆
(L⊙)
Lacc Ref. (Lacc)a
(L⊙)
R⋆
(R⊙)
Log Macc
(M⊙ yr−1)
M⋆
(M⊙)
Age
(Myr)
AA Tau
K7
CI Tau
K7
CW Tau
K3
CX Tau
M2.5
CY Tau
M1
DE Tau
M2
DL Tau
K7
DM Tau
M1
DN Tau
M0
DO Tau
M0
DR Tau
K7
DS Tau
K2
FM Tau
M0
FZ Tau
M0
GM Aur
K3
GO Tau
M0
HO Tau
M0.5
IQ Tau
M0.5
RY Tau
K1
SU Aur
G2
UZ Tau E M1
4060
4060
4730
3488
3705
3560
4060
3705
3850
3850
4060
4900
3850
3850
4730
3850
3778
3778
5080
5860
3705
0.80
1.01
1.32
0.48
0.55
1.01
0.89
0.24
0.91
0.79
1.09
0.90
0.26
1.06
1.26
0.27
0.16
0.73
6.59
10.6
2.27
0.03
0.11
0.05
0.03
0.04
0.07
0.23
0.10
0.02
0.60
0.72
0.21
0.10
0.46
0.07
0.01
0.01
0.04
1.60
0.55
0.31
1
2
2
2
1
1
2
2
1
1
2
1
2
3
1
3
2
2
4
4
2
1.8
2.0
1.7
1.9
1.8
2.6
1.9
1.2
2.2
2.0
2.1
1.3
1.2
2.3
1.7
1.2
0.9
2.0
3.3
3.2
3.7
-8.56
-7.95
-8.58
-8.18
-8.24
-7.69
-7.66
-7.99
-8.55
-7.10
-7.12
-8.00
-8.13
-7.14
-8.44
-9.14
-9.13
-8.24
-7.03
-7.52
-7.02
0.8
0.8
1.2
0.3
0.5
0.4
0.8
0.5
0.6
0.6
0.8
1.1
0.6
0.6
1.2
0.6
0.5
0.6
2.2
1.7
0.5
2.4
1.7
6.6
0.8
1.4
0.2
2.2
3.6
1.1
1.3
1.5
17.0
6.3
0.8
7.2
4.8
8.7
1.2
1.1
2.2
0.1
a) References. (1) Gullbring et al. (1998); (2) Valenti et al. (1993), these accretion luminosities have been multiplied by a factor
(140 pc/160 pc)2 to account for the current estimate of 140 pc for the distance to the Taurus-Auriga star forming region; (3)
Herczeg et al. (2009); (4) Calvet et al. (2004).
models for all sources are equal or less than the measured
flux at 7mm. Simultaneous 7mm and cm-wavelength obser-
vations with the EVLA will extend the coverage of the disk
SED to longer wavelengths, in order to better constrain the
millimeter-wavelengths dust opacity and to provide infor-
mation on dust grains larger than ∼ 1 cm.
3. Disk models
We compare the observational sub-mm/mm data with the
two-layer models of flared disks (i.e. in hydrostatic equi-
librium) heated by the stellar radiation as developed by
Dullemond et al. (2001), following the schematization of
Chiang & Goldreich (1997). These models have been used
in the analysis of CQ Tau by Testi et al. (2003) and of a
sample of 9 Herbig Ae stars by Natta et al. (2004) and we
refer to these papers for a more detailed description.
To completely characterize a model of the disk, we need
to specify the stellar properties (i.e. the luminosity L⋆, ef-
fective temperature Teff , mass M⋆ and distance d, assumed
to be 140 pc for all the sources in our sample), some char-
acteristics of the disk structure, namely the inner and outer
radius (Rin and Rout, respectively), the parameters Σ1 and
p defining the radial profile of the dust surface density as-
sumed to be a power-law4 Σdust = Σdust,1(R/1 AU)−p, the
disk inclination angle i (90o for an edge-on disk) and, lastly,
the dust opacity. The millimeter SED is almost completely
insensitive to some of the disk parameters, which cannot
thus be constrained by our analysis. For example, the inner
radius of the disk, Rin, affects the emission only at near
and mid-infrared wavelengths.
The outer radius Rout is a parameter that can be probed
by high-angular resolution interferometric observations. In
our sample 13 disks have been spatially resolved, yielding an
estimate for Rout. It is important to note that different in-
terferometric observations performed on the same disk can
lead to very different values of Rout (≥ 100 AU) because
4 As shown by Hughes et al. (2008) and Isella et al. (2009),
a tapered exponential edge in the surface density profile, physi-
cally motivated by the viscous evolution of the disk, is able to ex-
plain the apparent discrepancy between gas and dust outer radii
derived from millimeter observations of protoplanetary disks.
Here we want to note that when fitting the sub/mm-mm SED
with an exponential tail instead of a truncated power-law, the
dust opacity spectral index β (defined in Section 3.1) is un-
changed, and disk masses vary by less than a factor of 2, even
if the outer disk radius becomes much bigger (see, e.g., Table
4 in Kitamura et al. 2002). So the simpler truncated power-law
model is a justified approximation for the purposes of this paper.
Table 3. Literature sources for the sub-mm/mm data.
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
7
λ (mm)
Source Referencesa
0.450
0.600
0.624
0.769
0.800
0.850
0.880
1.056
1.100
1.200
1.300
1.330
2.000
2.700
3.100
3.400
7.000
1
2, 3
4
4
2, 3
1, 5
5
4
2, 3
6
1, 7, 8, 9, 10
5
2, 11
7, 8, 12, 13
8
9, 14
15
a) References. (1) Andrews & Williams (2005); (2) Mannings & Emerson (1994); (3) Adams et al. (1990); (4) Beckwith &
Sargent (1991); (5) Andrews & Williams (2007); (6) Althenoff et al. (1994); (7) Dutrey et al. (1996); (8) Jensen et al. (1996); (9)
Koerner et al. (1995); (10) Isella et al. (2009); (11) Kitamura et al. (2002); (12) Mundy et al. (1996); (13) Schaefer et al. (2009);
(14) Ohashi et al. (1996); (15) Rodmann et al. (2006).
of differences in the observations angular resolution, sen-
sitivity, and methods for deriving the radius. Given these
uncertainties in the determination of Rout, we consider a
range of possible values for Rout rather than a single-value
estimate, as listed in Table 4. For the 8 objects in our sam-
ple that have not yet been mapped with angular resolu-
tions smaller than a few arcseconds which may potentially
resolve the disk, a fiducial interval of 100 − 300 AU has
been assumed5. This adopted range for Rout comprises the
vast majority of the disk outer radii distribution function
as derived by Andrews & Williams (2007) through ∼ 1′′ an-
gular resolution sub-millimeter observations of 24 disks in
Taurus-Auriga and ρ−Ophiucus. We will discuss the impact
of the outer disk uncertainty on the results of our analysis
in Section 4. Here we want to note that with these values of
Rout the disk emission at sub-mm/mm wavelengths is dom-
inated by the outer disk regions (R > 30 AU) which are op-
tically thin to their own radiation. An effect of this is that
the millimeter SED does not depend on the disk inclina-
tion angle i (for i <∼ 80◦), thus decreasing the number of the
model parameters used to fit the observational SED. This is
certainly true for the spatially resolved disks for which we
have an estimate for Rout. The 8 unmapped objects could
5 We will refer to these sources (CW, CX, DE, DS, FM, FZ,
HO Tau, and SU Aur) as "unmapped" for the rest of the paper.
Note that they are all faint with F1mm < 120 mJy. This probably
explains why for these objects no observations at high-angular
resolution have been attempted yet: the sensitivity of the current
(sub-)mm interferometers would be not high enough to detect
the low surface brightness of the outer disk regions.
have optically thick inner disks that contribute significantly
to the sub-mm/mm emission. Since the disk outer radius
cannot be constrained by the SED alone, interferometric
imaging that spatially resolves these disks is the only way
to get information on the exact value of Rout. Nevertheless,
since the disks that have been spatially resolved so far show
outer radii typically in the range 100 − 300 AU, we consider
highly unlikely that a significant fraction of the unmapped
disks have Rout <
∼ 50 AU, which would make the disk opti-
cally thick even at mm wavelengths.
3.1. Dust opacity
The last quantity that defines the disk model is the dust
opacity, which depends on grain sizes, chemical composi-
tions, and shapes (e.g. Miyake & Nakagawa 1993, Pollack
et al. 1994, Draine et al. 2006). In order to extract from
the sub-mm/mm SED quantitative estimates for the dust
grain sizes and the dust mass in the disks, assumptions on
the chemical composition and shape of the dust grains have
to be made.
For our discussion we considered porous composite
spherical grains made of astronomical silicates, carbona-
ceous materials and water ice (optical constants for the
individual components from Weingartner & Draine 2001,
Zubko et al. 1996, Warren 1984 respectively) adopting frac-
tional abundances used by Pollack et al. (1994) (see caption
of Figure 3). We considered a population of grains with a
power-law size distribution n(a) ∝ a−q between a mini-
mum and a maximum size, amin and amax, respectively. In
8
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
the ISM, amin is a few tens of A, amax ∼ 0.1 − 0.2 µm,
and q = 3.5 (Mathis et al. 1977, Draine & Lee 1984, by
analysing extinction and scattering of starlinght from the
interstellar dust). As discussed e.g. in Natta et al. (2004)
in a protoplanetary disk, because of grain processing, one
expects much larger values of amax than in the ISM, and
a variety of possible q values, depending on the processes
of grain coagulation and fragmentation that occur in the
disk. Experimental studies of fragmentation from a single
target have found values of q ranging from ≈ 1.9 for low-
velocity collisions to ≈ 4 for catastrophic impacts (Davies
& Ryan 1990), but the power-law index for fragments from
shattering events integrated over a range of target masses
need not be necessarily the same as in the case of a single
target. In the disk, if the coagulation processes from the
initial ISM-grains population dominate over the fragmen-
tation ones, then a value for q that is lower than the ISM
one is expected (the opposite should occur in the case of
fragmentation playing the major role).
In the adopted two-layer disk model, the disk is divided
into the surface and midplane regions, and therefore two
different opacity laws are considered. Because of the as-
sumed vertical hydrodynamical equilibrium and dust set-
tling, in a protoplanetary disk the content of mass in the
midplane is much higher than in the surface. At millimeter
wavelengths, where the disk is mostly optically thin to its
own radiation, the midplane total emission dominates over
the surface one (although the surface dust plays a crucial
role for the heating of the disk). As a consequence, from our
analysis we can extract information only on the dust opac-
ity of the midplane, and for the rest of the paper when we
consider the dust opacity we will mean only the midplane
component.
For the dust grains in the disk surface we assumed the
same chemical composition and shape as for the midplane
grains, whereas for their size distribution we assumed for
amin and amax values around 0.1 µm (the sub-mm SED is
insensitive to these values as long as amax in the surface is
much lower than amax in the midplane, as expected if dust
settling is important).
Once the chemical composition and shape of the dust
grains in the midplane are set, the dust opacity law depends
on amin, amax and q. Considering a value of 0.1 µm for amin
(the dependence of the millimeter dust opacity from this
parameter turns out to be very weak), the models param-
eters for our analysis are the stellar parameters (L⋆, Teff,
M⋆), that have been set and listed in Table 2, plus the disk
ones (Σdust,1, p, Rout, q, amax), or equivalently (Mdust, p,
q, amax, Rout) where Mdust is the mass of dust in the disk.
For reasons described in Section 4 it is convenient to
approximate the millimeter dust opacity law discussed so
far in terms of a power-law at millimiter wavelengths κ =
κ1mm(λ/1 mm)−β, with κ1mm in units of cm2 per gram of
dust. At a fixed value of q the relations between κ1mm, β
and amax can be determined. These are shown in Figure 3
for q = 2.5, 3, 3.5, 4.
4. Results
The sub-mm/mm SED fits for all the objects in our sample
are reported in Figure 46. As discussed in Testi et al. (2003)
6 We did not include existing NIR and mid-IR data in our
analysis because at these shorter wavelengths the disk emission
and Natta et al. (2004), the only quantities that can be con-
strained by the sub-mm/mm SED are the millimeter dust
opacity spectral index β (in this paper calculated between
1 and 3 mm) and the Mdust × κ1mm product7. From the
SED alone it is impossible to constrain the value of the
surface density exponent p, since it is generally possible to
fit the SED data with either very flat (p = 0.5) or very
steep (p = 2) surface density profiles. However, all avail-
able high angular resolution observations performed so far
suggest that p ≤ 1.5 (e.g. Dutrey et al. 1996, Wilner et
al. 2000, Kitamura et al. 2002, Testi et al. 2003, Andrews
& Williams 2007, Isella et al. 2009, Andrews et al. 2009).
The degeneracy of the SED on p and Rout has an im-
pact onto the derived uncertainties for β and Mdust ×κ1mm.
Considering both this degeneracy and uncertainties of the
observational data, the total absolute uncertainties are ap-
proximately 0.2 − 0.3 for β and a factor of ≈ 3 and ≈ 4
for Mdust × κ1mm for the spatially resolved and unmapped
disks respectively. These uncertainties are mainly due to
the adopted ranges for Rout (listed in Table 4) and p be-
tween 0.5 and 1.5, whose values determine the impact of
the optically thick regions of the disk to the total emission,
as explained in Section 3.
The estimates for β and Mdust × κ1mm obtained by our
analysis are reported in Table 5.
4.1. Spectral slopes and dust opacity index
Before analysing the values of β, which provide information
on the level of grain growth in the outer disk regions, it is
important to check whether our results could be affected
by detection biases: indeed, for a given flux at ∼ 1mm, a
disk with large dust grains in the outer parts will produce a
stronger emission at ∼ 3 mm than a disk with a population
of smaller-sized grains and be easier to detect. In Figure 5
the grey region of the plot indicates the area in the (F1mm,
α1−3mm) plane in which the observations carried out so far
are not sensitive, due to the sensitivity limits obtained at 1
and 3 mm. While we could investigate the dust properties
for 4 disks with 15 mJy < F1mm < 30 mJy, the plot shows
that a disk with a 1 mm flux in this same range, but with a
spectral index α1−3mm >
∼ 3.2, consistent with a disk model
with ISM-like dust and an outer radius of ∼ 100 AU, would
not be detected at 3 mm. Nevertheless, we want to point
out that only 2 of our 16 targets with F1mm < 100 mJy
were not detected at 3mm, and both have very weak F1.3mm
fluxes (upper limits of 27 mJy and 21 mJy for DP Tau and
GK Tau respectively, Andrews & Williams 2005), therefore
the lack of disks with α ∼ 3.2 at F1mm > 30 mJy does not
seem to be due to our sensitivity limit but is most likely
real. Furthermore, as evident from the same plot, the 6
faintest disks, with F1mm < 60 mJy, show a spectral index
between 1 and 3 mm that is lower on average (2.08 ± 0.08)
than for the brighter disks (2.40 ± 0.06). This is the reason
of the discrepancy of the average value of the dust opacity
is optically thick and thus depends on the properties of the dust
grains located in the surface layers of the inner disk regions. The
aim of our analysis is instead to probe the optically thin disk
emission in order to investigate the dust properties in the disk
midplane.
7 Note that, according to the adopted disk model, after fix-
ing the stellar parameters, and these two quantities from the
observed SED, the midplane temperature is well constrained.
Table 4. Adopted disks outer radii.
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
9
Object
AA Tau
CI Tau
CW Tau
CX Tau
CY Tau
DE Tau
DL Tau
DM Tau
DN Tau
DO Tau
DR Tau
DS Tau
FM Tau
FZ Tau
GM Aur
GO Tau
HO Tau
IQ Tau
RY Tau
SU Aur
UZ Tau E
Ra
out (K02) Ra
(AU)
g (I09) Adopted Rout−interval
(AU)
(AU)
out (AW07) Rb
(AU)
214+63
−51
...
...
...
211 ± 28
...
152+28
−27
220 ± 47
147 ± 469
98 ± 52
193 ± 141
...
...
...
151+63
−29
...
...
329+403
−70
81 ± 128
...
...
400+600
−75
225 ± 50
...
...
...
...
175+50
−25
150+250
−100
100+300
−25
...
100+175
−25
...
...
...
150 ± 25
350+650
−175
...
...
150 ± 25
...
...
...
...
...
...
230
...
...
160
125
...
90
...
...
...
270
160
...
...
115
...
160
200 − 400
150 − 250
100 − 300
100 − 300
230 − 330
100 − 300
150 − 250
160 − 260
125 − 225
100 − 200
90 − 190
100 − 300
100 − 300
100 − 300
270 − 370
160 − 260
100 − 300
250 − 400
115 − 215
100 − 300
160 − 260
a) Rout listed here results from image fitting using a truncated power-law for the surface density profile. K02: Kitamura et
al. (2002); AW07: Andrews & Williams (2007). b) Rg is the radius within which the 95% of the source flux is observed, after
fitting the visibility of the source with a gaussian; we used this value, when available, as a lower limit for our adopted
Rout-interval. I09: Isella et al. (2009).
spectral index < β > between unmapped and resolved disks
(see discussion below).
In Figure 3 we plot κ1mm and β as functions of amax
for different values of q. One interesting result is that the
values of β obtained for the objects in our sample by the
SED-fitting procedure are all less than or equal to 1, which
can be explained only if dust grains with sizes larger than 1
mm are present in the disk. For the 9 sources with β ≤ 0.5,
the largest dust grains must have grown to sizes ≥ 1 cm
according to our dust model. If we let q assume values lower
than 2.5, thus increasing the fractional number of the larger
grains, the maximum grain size decreases, but amax >∼ 1 mm
for β <∼ 1 is still needed.
The estimates of the maximum grain sizes amax depend
on the dust model that one adopts. If we consider for ex-
ample the dust model adopted in Natta et al. (2004) (5%
olivine, 15% organic materials, 35% water ice and 50% vac-
uum), the estimates for amax are greater by about an order
of magnitude than the ones obtained with our dust model.
In general, no realistic dust grain models result in β < 1
for amax < 1 mm.
Here it is worthwhile to remember that, whereas for the
13 resolved disks in our sample β is well determined, for the
8 disks for which we do not have information on their spatial
extension, the optically thick parts of a possible compact
disk could significantly contribute to the emission. If that is
the case, no conclusions on the dust grain properties could
be drawn. In particular, for the 6 unmapped disks with
α ≈ 1.9 − 2.2 (CX Tau, DS Tau, FM Tau, FZ Tau, HO Tau,
SU Aur) a very compact disk with Rout ≈ 20 AU, a relative
high inclination i ≈ 70o, and with ISM-like dust grains
would fit the sub-mm/mm SED. For the other 2 unmapped
sources, DE Tau (α ≈ 2.4) and CW Tau (α ≈ 2.5), ISM-
like dust grains are consistent with the sub-mm/mm data
using a disk model with i ≈ 70o and Rout ≈ 30, 40 AU
respectively. For the faintest disks, the outer radius has to
be small because if the disks were larger they would also
need to be more massive in order to keep being optically
thick, thus producing an α value close to 2 with all the
possible dust properties. In that case, they would emit more
sub-mm/mm-wave radiation than observed. Since, for an
optically thick disk, decreasing the inclination angle has
the effect of making the SED steeper (because the central,
hotter parts are less obscured), for disks with i < 70o the
outer radii would need to be even smaller to fit the SED.
Since these disk outer radii are lower by factors of 2 − 4
10
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
Fig. 3. Millimeter dust opacity for the adopted dust model. Left top panel: dust opacity at 1 mm as function of amax for
a grain size distribution n(a) ∝ a−q between amin = 0.1 µm and amax. Different curves are for different values of q, as
labelled. The dust grains adopted here are porous composite spheres made of astronomical silicates (≈ 10% in volume),
carbonaceous materials (≈ 20%) and water ice (≈ 30%; see text for the references to the optical constants). Left bottom
panel: β between 1 and 3 mm as a function of amax for the same grain distributions. Right panel: dust opacity at 1 mm
as function of β between 1 and 3 mm for the same grain distributions; different iso-amax curves for amax = 0.1, 1, 10 cm
are shown. In the plots the unit of measure for the dust opacity is cm2 per gram of dust.
the unmmapped sources are compact, optically-thick disks
with unprocessed ISM grains. For AA Tau, which is the
faintest of the mapped sources (F1mm ≈ 100 mJy) and has
a low spectral index of α ≈ 2.0 typical of fainter sources,
the maps indicate a disk radius of 200 − 400 AU ruling out
the possibility of optically thick disk regions dominating
the total emission, and requiring a low β of 0.3 to explain
the shallow slope.
∼ 0.8, values of q >
Also, only for the disks with β >
∼ 3.5
are consistent with the data (see right panel in Figure 3), in
which case amax ≥ 10 cm. This is in good agreement with
the following consideration using the approximated asyn-
thotic formula β ≈ (q − 3)βRJ , derived by Draine (2006),
where βRJ is the opacity spectral index of the solid material
in the Rayleigh limit, and valid for a population of grains
with 3 < q < 4 and amin ≪ ac ≪ amax, with ac being a
function of the real and imaginary parts of the complex di-
electric function and wavelength (at millimeter wavelengths
ac ∼ 1 mm). Since, at a fixed q, the function β(amax) de-
creases when amax >∼ 1 mm (see Figure 3), it follows that
the approximated asynthotic value (q − 3)βRJ gives a lower
limit for β at a finite amax. Adopting βRJ = 1.7 from the
small dust grains in the ISM, for q >∼ 3.5 we derive β >∼ 0.85,
i.e. dust grains populations with q >
∼ 3.5 cannot explain β-
values lower than ≈ 0.858.
In Figure 6 we report the histogram of the β-values
derived by our analysis. The average value in our sample
considering both unmapped and resolved disks, < β >=
0.6 ± 0.06, is not consistent with dust populations with q-
values of 3.5, like for the ISM, or greater. Values of q lower
than 3.5 have been obtained by, e.g., Tanaka et al. (2005)
from simulations of grain growth in stratified protoplane-
tary disks which give vertically integrated size distributions
8 Note that β-values greater than ≈ 0.85 can be explained by
both q <∼ 3.5 and q >∼ 3.5.
Fig. 5. Spectral
index α between 1 and 3 mm plotted
against the flux at 1 mm from the model best fits shown
in Figure 4. The points plotted here are listed in the first
two columns of Table 5. Open blue points represent the
spatially resolved disks, whereas the filled red points are
for the unmapped ones. A value of α consistent with a disk
model with ISM-like dust and an outer radius of ∼ 100 AU
is indicated as a dashed horizontal line. The grey region
shows the area in the plot that is unexplored because of
the 3σ sensitivity limits of our new and past millimeter ob-
servations (here considered as 15 mJy and 1.0 mJy at 1 and
3 mm respectively).
than the smallest disks that have been spatially resolved
so far, and since for all the 13 resolved disks the values
of β have been found always much lower than in the ISM
(βISM ≈ 1.7, see Figure 6), thus showing clear evidence
of grain growth, we consider very unlikely that many of
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
11
Fig. 4. Fits of the sub-mm/mm SEDs with the two-layer flared disk models (solid lines). Note that even if the data at 7
mm, available only for 8 disks in our sample, are present in the plots, they have not been used in the fitting procedure
(see Section 2.5). The errorbars in the plots take into account an uncertainty of 10% on the absolute flux scale, typical
for flux measurements at these wavelengths. The circled data come from the new PdBI observations. The fitting values
of the spectral index α and of the dust opacity spectral index β between 1 and 3 mm are indicated in the bottom left
corner of each plot. The value of the surface density power-law index p adopted in these fits is 1, and the adopted value
for the outer disk radius is listed in Table 5.
with q ≈ 3, plus a population of much larger bodies, to
which however our observations would not be sensitive.
These considerations clearly suggest that the process of
dust grain growth to mm-sizes has played a fundamental
role in most, if not all, of the Taurus-Auriga protoplane-
tary disks whose sub-mm/mm SED has been sufficiently
investigated.
From Figure 6 it is also evident that the two sub-samples
made of unmapped and resolved disks separately have dif-
ferent averaged β values, with the β values being on av-
erage lower for the unmapped disks than for the resolved
ones (< β >= 0.41 ± 0.10, 0.68 ± 0.06 for the unmapped
and resolved disks respectively). Since the unmapped sam-
ple nearly coincides with the fainter disks with F1mm < 100
mJy, and since β is correlated with the spectral index α,
this is due to the trend already presented in Figure 5 be-
tween F1mm and α1−3mm. This trend agrees with simple
expectations for grain growth: as the maximum grain size
12
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
Table 5. Disk properties.
Object namea
Rb
out
(AU)
F c
1mm
(mJy)
α1−3mm
β1−3mm Mdust × κ1mm
(M⊙× cm2g−1)
M q=2.5
dust
(M⊙)
M q=3
dust
(M⊙)
M q=3.5
dust
(M⊙)
AA Tau
CI Tau
CW Tau
CX Tau
CY Tau
DE Tau
DL Tau
DM Tau
DN Tau
DO Tau
DR Tau
DS Tau
FM Tau
FZ Tau
GM Aur
GO Tau
HO Tau
IQ Tau
RY Tau
SU Aur
UZ Tau E
300
200
200
200
280
200
200
210
175
150
140
200
200
200
320
210
200
325
165
200
210
108
314
129
19
168
69
313
209
153
220
298
28
29
27
423
151
36
118
383
50
333
2.0
2.5
2.5
2.3
2.2
2.5
2.4
2.5
2.3
2.2
2.7
2.0
1.9
2.1
2.6
2.6
2.1
2.2
2.3
2.1
2.5
0.3
0.8
0.8
0.5
0.5
0.7
0.7
1.0
0.5
0.4
1.0
0.2
0.2
0.2
1.0
1.0
0.5
0.5
0.5
0.2
0.7
2.8 · 10−4
7.9 · 10−4
3.2 · 10−4
5.7 · 10−5
6.0 · 10−4
1.6 · 10−4
8.4 · 10−4
9.2 · 10−4
3.5 · 10−4
4.9 · 10−4
6.6 · 10−4
6.2 · 10−5
6.5 · 10−5
5.0 · 10−5
1.4 · 10−3
6.3 · 10−4
1.7 · 10−4
3.4 · 10−4
5.1 · 10−4
5.0 · 10−5
6.1 · 10−4
1.5 · 10−4
1.1 · 10−4
4.2 · 10−5
1.2 · 10−5
1.3 · 10−4
2.4 · 10−5
1.3 · 10−4
1.0 · 10−4
7.7 · 10−5
1.4 · 10−4
7.4 · 10−5
4.0 · 10−5
6.0 · 10−5
3.2 · 10−5
1.6 · 10−4
7.1 · 10−5
3.7 · 10−5
7.4 · 10−5
1.1 · 10−4
3.2 · 10−5
9.6 · 10−5
1.5 · 10−3
1.2 · 10−4
4.4 · 10−5
1.9 · 10−5
2.0 · 10−4
2.8 · 10−5
1.5 · 10−4
1.0 · 10−4
1.2 · 10−4
3.1 · 10−4
7.2 · 10−5
1.1 · 10−3
1.7 · 10−3
9.0 · 10−4
1.5 · 10−4
6.9 · 10−5
5.6 · 10−5
1.1 · 10−4
1.7 · 10−4
9.0 · 10−4
1.1 · 10−4
...
5.0 · 10−4
1.9 · 10−4
...
...
...
...
1.6 · 10−4
...
...
1.2 · 10−4
...
...
...
2.5 · 10−4
1.1 · 10−4
...
...
...
...
...
a) Underlined objects are those that have not been mapped to date through high-angular resolution imaging. b) Central value of
the adopted Rout-interval reported in Table 4. This is the value used to extract the other quantities listed in the Table. c) Source
flux density at 1 mm from the best fit disk model described in Section 3.
amax increases, both β and κ1mm decrease, so disks with the
same dust mass should become fainter as β gets smaller.
Alternatively, the F1mm-α1−3mm trend could arise if the
fainter disks are smaller, and then more optically thick,
than the brighter, resolved ones, while their "true" β would
be nearly the same.
In Figure 7 we show the β vs stellar age plot. No ev-
idence of an evolutionary trend is present over the large
range of stellar ages spanned by our sample. As already
discussed the determination of the individual ages of PMS
stars is very uncertain. Hartmann (2001) has shown that
the large age spread deduced for T Tauri stars in Taurus
may just be the result of uncertainties in the measurements
towards individual members of a population with a very
narrow age spread. Our most solid conclusion, not affected
by the age determination uncertainties, is that most (if not
all) of the Class II YSO in our sample show evidence of
evolved grains in the disk outer regions.
4.2. Grain growth
These results confirm two statements that have been ad-
dressed in the last years. The first one is that the dust
coagulation from ISM dust to mm/cm-sized grains is a fast
process, that appears to occur before a YSO enters in the
Class II evolutionary stage, i.e. when a dense envelope still
surrounds the protostar+disk system. This is also consis-
tent with the recent results by Kwon et al. (2009) who have
found evidence of grain growth in 3 Class 0 YSOs in Perseus
and Cepheus through CARMA observations at 1.3 and 2.7
mm.
The second statement is that the mm/cm-sized dust
grains appear to stay longer in the outer disk if com-
pared with their expected inward drift timescale due to
the interaction with the gas component. In a circum-
stellar disk the gas experiences a pressure gradient force
which cancels a part of the central star's gravity and in-
duces a rotation slower than the Keplerian velocity. Dust
grains, whose orbits are hardly affected by gas pressure,
are assumed to be rotating with Keplerian velocity, and
thus experience a headwind. They lose angular momentum
and spiral inward to the central star (Adachi et al. 1976,
Weidenschilling 1977). Takeuchi & Lin (2005) proposed as
possible solutions to the drift problem that either the grain
growth process from mm-sized to cm-sized grains is very
long in the outer regions of the disk (timescales greater than
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
13
cal orbital decay time for a cm-sized grain with a density of
0.1 g cm−3 at 50 AU is ≈ 105 yr (104 yr for a more compact
grain with a 1 g cm−3 density), to explain the presence of
cm-sized grains in the older disks in our sample (with ages
of few Myrs and up to ∼ 107 yr) the reservoir of larger bod-
ies would have to contain at least 100 times more mass than
the observed dust mass. The total disk mass would then be
comparable to or even exceeding the estimated stellar mass
(using a standard gas-to-dust mass ratio of 100, see Section
4.3).
To solve the drift problem for very low and very high
disk masses Brauer et al. (2007) proposed a significant re-
duction of the drag experienced by the dust due to a reduc-
tion of the gas-to-dust ratio in the disk from the canonical
value of 100, and to possible collective effects of dust gath-
ering in a thin midplane layer in the case of very-low tur-
bulent disks. For example, using a low value for the turbu-
lence parameter α = 10−6 and a gas-to-dust ratio of 5, they
find that cm-sized dust grains remain in the outer parts of
the disk for more than 2 Myr if disk masses < 0.05 M⋆ or
> 0.2 M⋆ are considered. In the range Mdisk ≈ 0.05−0.2 M⋆
the collective effects of dust are not efficient enough to re-
tain cm-sized grains in the disk outer regions. If the disk is
more turbulent, then more gas has to be removed in order
to keep the particles in the outer parts for a longer period
of time. To justify such small gas-to-dust ratios tha authors
suggest the process of photoevaporation as the main driver
of gas removal. However, in this case, since photoevapo-
ration is expected to manifest only in the later stages of
disk evolution, then we would expect a strong evolutionary
trend of β, which is instead not found by our analysis.
Another possibility is the presence in the disk of local
gas pressure maxima that would block the inward drift of
the dust grains. As shown by the dynamics calculation of
Barge & Sommeria (1995) and Klahr & Henning (1997) this
mechanism may provide a solution also to the relative veloc-
ity fragmentation barrier that prevents meter-sized bodies
to further grow in size in the inner regions of the disk.
4.3. Dust mass
From the sub-mm/mm SED fitting procedure we derived
the quantity Mdust × κ1mm. To obtain the dust mass in
the disks an estimate of κ1mm is needed. The choice of the
dust opacity at 1 mm creates the largest uncertainty in the
derivation of (dust) disk masses from millimeter observa-
tions. Estimates of κ1mm from different dust models in the
literature span a range of a factor greater than 10. The
most commonly used value in the literature is κ1mm ≃ 3
cm2 per gram of dust (Beckwith et al. 1990) that, for sim-
plicity, is supposed to be constant, i.e. independent of the
dust grain properties, for all the disks. As noted in Natta
et al. (2004) this is not realistic as grain growth affects
both the frequency dependence of the dust opacity and its
absolute value (e.g. Miyake & Nakagawa 1993).
In this paper we use the dust model described in Section
3.1 to get an estimate of κ1mm that takes into account the
dust physical properties, expecially grain growth. As shown
in the left top panel of Figure 3, at a fixed value of the grain
size distribution index q, the value of κ1mm can be inferred
from amax which in turn can be derived from β through the
relation plotted in the left bottom panel of the same figure.
The relation between κ1mm and β is shown in the right
panel. Therefore, from an estimate of β and Mdust × κ1mm
Fig. 6. Distribution of the dust opacity spectral index β for
the disks in our sample. In blue the values for the spatially
resolved disks are indicated, whereas the red is for the un-
mapped ones. The average β value for all the sources in
our sample (black), for the unmapped sources only (red),
for the resolved sources only (blue), and the value of 1.7 for
the ISM dust are indicated as dashed vertical lines.
Fig. 7. Beta versus age: relationship between the dust opac-
ity spectral index β and the estimated stellar age obtained
as in Section 2.3. Open blue points represent the spatially
resolved disks, whereas the filled red points are for the un-
mapped ones. The β value for the ISM dust is indicated as
a dashed horizontal line.
1 Myr, e.g. given by grains' sticking probability as low as
0.1) or that grains have already grown to 10 m or larger and
they do not migrate rapidly anymore, and mm-sized parti-
cles are continuosly replenished through collisions of these
large bodies. However our analysis suggests that, unless the
grain size distribution index is very low, i.e. q < 2.5, disks
with β <
∼ 0.5 show evidence of grains as large as ∼ 1 cm in
the outer regions. Also, as already pointed out in Natta et
al. (2004) and Brauer et al. (2007), a noteworthy amount of
larger bodies, which would not contribute to the disk emis-
sion because of their negligibly small opacity coefficient,
could significantly increase the disk masses. Since the typi-
14
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
obtained by the sub-mm/mm SED fitting procedure we can
get an estimate for κ1mm and then the dust mass Mdust for
our dust model, after fixing the q parameter. In Table 5 we
report the values of Mdust as obtained with this method for
q = 2.5, 3.0, 3.5. It is important to bear in mind that, as
for the estimate of the maximum grain size, also for κ1mm,
and thus for the dust mass, the estimates depend on the
specific dust model that one adopts (see discussion in Natta
et al. 2004). However, following Isella et al. (2009), we take
into account the variation of the dust opacity with β using
a well defined dust model to infer dust masses, and this
approach is more accurate than considering a fixed κ1mm
as often done in previous studies.
dust
dust
dust
dust < M q=3
dust < M q=3.5
dust
dust
and M q=3
and M q=3
is greater than M q=2.5
In general, for our specific dust model, when β < 1
then M q=2.5
, the amount of variation
of M q
dust for the different q values being dependent on β.
In particular, when β ≃ 0.8 − 1 M q=2.5
dust are
nearly coincident and M q=3.5
by
factors of ∼ 2 − 4. For β <∼ 0.7 a grain size distribution
with q = 3.5 cannot explain the observational data and the
discrepancy between M q=2.5
dust increases up to a
factor of ∼ 15 when β ≃ 0.2. This is due to the increase of
the distance between the κ1mm(β) curves at fixed q-values,
when β decreases, as shown in the right panel of Figure 3
As already mentioned in Section 4.1, from the value of
β it is impossible to disentangle amax and q. Nevertheless,
only the disks showing β ≥ 0.8 can be explained by a grain
size distribution index q = 3.5. For this reason, for the rest
of our analysis we will consider separately only the cases
in which the dust grain size distribution is characterized
by q = 2.5 or 3. The value of q in protoplanetary disks is
unknown and it is not observationally constrainable, and
so it is in principle possible that it can vary substantially
from one disk to the other. For our discussion we will use
the assumption that q is the same for all the disks.
Once the dust mass is estimated, this is commonly con-
verted into an estimate of the total disk mass (dust+gas)
using the standard ISM gas-to-dust ratio of 100. However,
if the dust and gas components have different evolutionary
timescales, this ratio is a function of time and the stan-
dard ISM value would be inappropriate for estimating the
disk mass at the present time. Physical processes that are
likely to alter this ratio during the evolution of a proto-
planetary disk are gas photo-evaporation, which would lead
to a decrease of the gas-to-dust ratio (unless a significant
amount of dust is dragged away by the gas flow), and dust
inward migration, which instead would increase it. Natta et
al. (2004) suggested that the standard ISM ratio provides a
correct estimate of the "original" disk mass, when the disk
composition reflected that of the parent cloud. However,
this is true only if a negligible amount of dust has migrated
with respect to the gas and accreted onto the central star,
otherwise the ISM ratio would provide only a lower limit
to the original disk mass. Because of our ignorance on the
gas-to-dust ratio evolution in this paper we will not convert
the dust mass into a total disk mass. Also, it is important
to note that since millimeter observations are insensitive
to the emission of bodies much larger than 1 cm, including
planetesimals and planets, the dust mass we are referring to
in this paper does not include the contribution from these
larger objects. The dust mass is intended to be the mass in
"small grains" only, and is a lower limit to the total mass
in dust.
Figure 8 shows the β vs Mdust plots for q equal to 2.5
and 3. The plot with q = 2.5 does not show any signifi-
cant relation. In the q = 3 case an anticorrelation between
these two quantities may be present, with the more massive
dust disks being associated to the ones showing today the
largest grains in their outer parts, with amax up to ∼ 10
cm. This possible anticorrelation is clearly an effect of the
choice of a physical model for the dust that takes into ac-
count the variation of the dust opacity normalization factor
with the size-distribution of the dust grains (Section 3.1).
If we had made the unrealistic assumption of a constant
value for the dust opacity at, say, 1 mm, we would have
found an opposite result, i.e. disks with less mass in dust
would have shown on average larger grains, as can be de-
rived from Figure 5, since α is approximately proportional
to β and F1mm would be approximately proportional to
the dust mass, in the constant κ1mm case. More realisti-
cally, lower values of β imply low absolute values of the
dust millimeter opacity (see right panel of Figure 3), so
that, for a fixed observed flux and disk temperature, the
corresponding dust mass must be larger. The reason why
the plot with q = 2.5 does not show signs of anticorrela-
tion is that at a fixed β a lower q value is associated to a
smaller dust grain maximum size. The β range spanned by
the disks in our sample translates into a range in κ1mm that
is narrower for q = 2.5 than it is for q = 3 (see right panel
of Figure 3), thus leading to smaller correction factors with
respect to the constant κ1mm case. Note that among the 5
disks showing the smallest β (≈ 0.2 − 0.3) and the highest
dust masses (about 10−3 M⊙ for q = 3) only one (AA Tau)
has been mapped and spatially resolved. High angular res-
olution and high sensitivity imaging of these disks together
with other ones are needed to test the robustness of the
β − Mdisk,init anticorrelation in the q = 3 case.
Figure 9 and 10 plot the disk mass in dust against the
estimated mass of the central star (dust mass obtained
with q-values of 2.5 and 3) and mass accretion rate (dust
mass with q = 3 and stellar ages smaller and greater
than 2 Myr) respectively. In the 0.3 − 2 M⊙ range in
stellar mass the dust mass does not show any trend (for
q = 2.5: 3 · 10−5 < Mdust < 2 · 10−4M⊙; for q = 2.5:
4 · 10−5 < Mdisk,init < 2 · 10−3M⊙). The two YSOs with
M⋆ <∼ 0.3 M⊙ have lower dust masses than all the other
YSOs for both q = 2.5 and 3. However, the range in stellar
mass spanned by our sample is rather limited, and partic-
ularly biased toward YSOs with late-K and early-M spec-
tral types. From the upper limits obtained at 1.3 mm by
Schaefer et al. (2009) for 14 T-Tauri stars with spectral
types later than ≈ M2 (F1.3mm <
∼ 5 − 20 mJy), and from the
detection of very low disk masses around sub-stellar objects
in Taurus-Auriga (Klein et al. 2003, Scholz et al. 2006), it is
likely that a correlation between the disk mass in dust and
the mass of the central star exists. High sensitivity millime-
ter observations are needed to constrain the disk masses in
dust around very low mass PMS stars. In Figure 10 a trend
of higher mass accretion rates for the more massive disks
in dust may be present in the data but still the scatter
is very large (a similar result is obtained for the q = 2.5
case). Other relations between the dust properties (β, dust
mass) and the stellar ones (e.g. stellar luminosity, effective
temperature) have been investigated but no significant cor-
relations have been found.
Finally, we have compared the values of β obtained in
this paper with the ones derived in Natta et al. (2004) for
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
15
Fig. 8. Dust opacity spectral index β and maximum grain size amax plotted against the disk mass in dust for the sources
in our sample. In the left and right panels the dust mass has been obtained adopting a value of 2.5 and 3 respectively for
the power index q of the dust grains size distribution. Open blue points represent the spatially resolved disks, whereas
the filled red points are for the unmapped ones. The errorbars in the dust mass estimates do not take into account our
uncertainty on the real value of q.
Fig. 9. Disk mass in dust plotted against the stellar mass for the sources in our sample. In the left and right panels
the dust mass has been obtained adopting a value of 2.5 and 3 respectively for the power index q of the dust grains
size distribution. The stellar mass has been estimated as described in Section 2.3, and the errorbars take into account
the range of values given by different PMS stars evolutionary models. Open blue points represent the spatially resolved
disks, whereas the filled red points are for the unmapped ones. The errorbars in the dust mass estimates do not take into
account our uncertainty on the real value of q.
a sample of 6 intermediate mass Herbig Ae stars and 3
lower mass T Tauri stars. From the two-sided Kolmogorov-
Smirnov test, the probability that the β values from the two
samples come from the same distribution is about 88%, in-
dicating that the disks in the two samples have similar dust
properties in terms of grain growth (in the Natta et al. 2004
sample only one object, the Herbig Ae star HD150193, has
a relatively high β = 1.6±0.2, compatible with non-evolved
ISM-like dust grains).
5. Summary
We have analysed the sub-mm/mm SED out to 3mm of a
sample of 21 Class II T-Tauri stars in the Taurus-Auriga
star forming region. This sample comprises about the 60%
of the "isolated" class II YSOs with an estimated stellar
mass greater than ≈ 0.4 M⊙. It also comprises approxi-
mately 1/3 of the isolated class II wich have been observed
at 0.85mm to have a flux less than 100 mJy, for which we
provide first 3mm detections thanks to sensitive PdBI ob-
servations. Our main findings are summarized below:
1. For all the sources in our sample the millimeter opacity
spectral index β is significantly lower than the value
16
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
Fig. 10. Disk mass in dust plotted against the mass accretion rate for the sources in our sample. In the left and right
panels we included the objects with estimated ages of less and more than 2 Myr respectively. The dust mass has been
obtained adopting a value of 3 for the power index q of the dust grains size distribution. The stellar mass accretion
rate has been estimated following the method described in Section 2.4. Open blue points represent the spatially resolved
disks, whereas the filled red points are for the unmapped ones. The errorbars in the mass accretion rate take into account
the typical intrinsic time-variability of the accretion rates. The errorbars in the dust mass estimates do not include our
uncertainty on the actual value of q.
obtained for the dust in the ISM. For the 13 spatially
resolved disks this is a clear evidence of the presence in
the outer disk regions of dust grains as large as at least 1
mm. For the 8 unmapped sources, assuming disk outer
radii comparable to the resolved sources low β values
are found as well, suggesting that even for these fainter
sources 1 mm-sized grains are in the disk outer regions.
This confirms, over a larger sample and less massive
disks, the results obtained by past observations on T-
Tauri stars (e.g. Rodmann et al. 2006 and references
therein).
2. No significant evidence of an evolutionary trend for the
millimeter opacity spectral index β has been found. This
indicates that the dust grain growth to ∼ 1 mm-sizes is a
very fast process in a protoplanetary disk, that appears
to occur before a YSO enters in the Class II evolutionary
stage. Also, the amount of these large grains in the disk
outer regions does not appear to decline throughout all
the Class II evolutionary stage.
3. Only for 9 sources in our sample a power law index for
the grain size distribution of the ISM, qISM ≈ 3.5, is
consistent with the sub/mm-mm SED. Instead, most of
the disks show evidence of a smaller q−value.
4. We have not found any significant correlation between
the dust properties, namely grain growth and dust mass
in the disk outer regions, and the properties of the cen-
tral star, including the mass accretion rate onto the stel-
lar surface. However, our sample contains mostly YSOs
with a narrow range in spectral types, namely between
late-K and early-M, and it is possible that a correlation
between the disk mass in dust and the stellar mass will
become apparent when considering also disks around
late-M spectral types.
5. The 6 faintest sources in our sample, with F1mm < 50
mJy, show a spectral index between 1 and 3 mm that is
on average lower than the spectral index of the brighter
sources. This may indicate either that for these fainter,
yet unmapped disks the emission from the optically
thick inner disk is much more significant than for the
brighter sources or that the dust grains in the outer
regions of these fainter disks are even larger than in
the brighter YSOs, possibly up to 10 cm if q = 3. The
latter hypothesis would imply on average a lower abso-
lute value for the millimeter opacity coefficient for these
fainter disks. Taking into account this effect through a
physical model for the dust properties, we found, in the
case of a value for the power index of the grain size
distribution q = 3, an anticorrelation between β and
the disk dust mass, that would indicate that larger dust
grains are found in the outer parts of the more massive
disks (in dust).
In the next future, with the advent of the next sub-
mm/mm facilities (both new interferometers like ALMA
and technical improvements in the existent arrays, like for
example the E-VLA) it will be possible to extend the in-
vestigation of the dust properties to fainter circumstellar
disks around the lowest mass PMS stars and brown dwarfs.
Furthermore, thanks to the great enhancements in both
high angular resolution and sensitivity that these facilities
will provide, it will be possible to determine their spatial
extension.
Acknowledgements. We thank the anonymous referee for his/her
comments that greatly improved the manuscript. L.R. acknowledges
the PhD fellowship of the International Max Planck Research School.
References
Adachi, I., Hayashi, C., & Nakazawa, K. 1976, Prog. Theor. Phys. 56,
1756
Adams, F. C., Emerson, J. P., & Fuller, G. A. 1990, ApJ 357, 606
Altenhoff, W. J., Thum, C., & Wendker, H. J. 1994, A&A 281, 161
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., & Dullemond,
C. P. 2009, ApJ 700, 1502
Andrews, S. M., & Williams, J. P. 2007, ApJ 659, 705
Ricci et al.: Protoplanetary Disks in Taurus-Auriga
17
Andrews, S. M., & Williams, J. P. 2005, ApJ 631, 1134
Baraffe, I., Chabrier, G., Allard, F. & Hauschildt, P. H. 1998, A&A
337, 403
Barge, P., & Sommeria, J. 1995, A&A 295, L1
Beckwith, S. V. W., Sargent, A. I., Chini, R. S., & Guesten, R. 1990,
AJ 99, 924
Beckwith, S. V. W., & Sargent, A. I. 1991, ApJ 381, 250
Bertout, C., Robichon, N., & Arenou, F. 1999, A&A 352, 574
Bessell, M. S., & Brett J. M. 1988, PASP 100, 1134
Brauer, F., Dullemond, C. P., Johansen, A. , et al. 2007, A&A 469,
1169
Bouwman, J., Meeus, G., de Koter, A., Hony, S., Dominik, C., &
Waters, L. B. F. M. 2001, A&A 375, 950
Briceno, C., Hartmann, L., Stauffer, J. R., & Martin, E. L. 1998, AJ
115, 2074
Padgett, D. L., Brandner, W., Stapelfeldt, K. R., Strom, S. E.,
Terebey, S., & Koerner, D. 1999, AJ 117, 1490
Palla, F., & Stahler, S. W. 1999, ApJ 525, 772
Pinte, C., Fouchet, L., M´enard, F., Gonzalez, J.-F., & Duchene, G.
2007, A&A 469, 963
Pollack, J. B., Hollenbach, D., Beckwith, S., Simonelli, D. P., Roush,
T., & Fong, W. 1994, ApJ 421, 615
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J., Podolak,
M., & Greenzweig, Y. 1996, Icarus 124, 62 Hollenbach, D.,
Beckwith, S., Simonelli, D. P., Roush, T., & Fong, W. 1994, ApJ
421, 615
Reipurth, B., & Zinnecker, H. 1993, A&A 278, 81
Rieke, G. H., & Lebofsky, M. J. 1985, ApJ 288, 618
Rodmann, J., Henning, T., Chandler, C. J., Mundy, L. G., & Wilner,
D. J. 2006, A&A 446, 211
Briceno, C., Luhman, K. L., Hartmann L., Stauffer, J. R., &
Kirkpatrick, J. D. 2002, ApJ 580, 317
Safronov, V. S., & Zvjagina, E. V. 1969, Icarus 10, 109
Schaefer, G. H., Dutrey, A., Guilloteau, S., Simon, M., & White, R.
Calvet, N., Muzerolle, J., Briceno, C., Hernndez, J., Hartmann, L.,
J. 2009, ApJ, 710, 698
Saucedo, J. L., & Gordon, K. D. 2004, AJ 128, 1294
Schmidt-Kaler T. 1982, in Landolt-BornsteinGroup VI, Vol. 2, ed.
Calvet, N., D'Alessio, P., Hartmann, L., Wilner, D., Walsh, A., &
K.-H. Hellwege (Berlin: Springer), 454
Scholz, A., Jayawardhana, R., & Wood, K. 2006, ApJ 645, 1498
Simon, M., & Prato, L. 1995, ApJ 450, 824
Simon, M., Chen, W.-P., Howell, R. R., Benson, J. A., & Slowick, D.
1992, ApJ 384, 212
Takeuchi, T., & Lin, D. N. C. 2005, ApJ 623, 482
Tanaka, H., Himeno, Y., & Ida, S. 2005, ApJ 625, 414
Testi, L., Natta, A., Shepherd, D. S., & Wilner, D. J. 2001, ApJ 554,
1087
Testi, L., Natta, A., Shepherd, D. S., & Wilner, D. J. 2003, A&A 403,
323
Throop, H. B., Bally, J., Esposito, L. W., & McCaughrean, M. J.
2001, Science 292, 1686
Valenti, J. A., Basri, G., & Johns, C. M. 1993, AJ 106, 2024
Warren, S. G. 1984, ApOpt 23, 1206
Weidenschilling, S. J. 1977, MNRAS 180, 57
Weingartner, J. C., & Draine, B. T. 2001, ApJ 548, 296
Wilner, D. J., Ho, P. T. P., Kastner, J. H., & Rodr´ıguez, L. F. 2000,
ApJ 534, L101
Zubko, V. G., Mennella, V., Colangeli, L., & Bussoletti, E. 1996,
MNRAS 282, 1321
Sitko, M. 2002, ApJ 568, 1008
Calvet, N., & Gullbring, E. 1998, ApJ 509, 802
Carpenter, J. M. 2001, AJ 121, 2851
Chiang, E. I., & Goldreich, P. 1997 ApJ 490, 368
Cutri, R. M. et al. 2003, 2MASS All Sky Catalog of point sources
Davies, D. R. & Ryan, E. V. 1990, Icarus 83, 156
Draine, B. T. & Lee, H.-M. 1984, ApJ 285, 89
Draine, B. T. 2006, ApJ 636, 1114
Dullemond, C. P., Dominik, C., & Natta, A. 2001, ApJ 560, 957
Dutrey, A., Guilloteau, S., Duvert, G., et al. 1996, A&A, 309, 493
Gullbring, E., Calvet, N., Muzerolle, J., & Hartmann, L. 2000, ApJ
544, 927
Gullbring, E., Hartmann, L., Briceno, C., & Calvet, N. 1998, ApJ 492,
323
Hartigan, P., & Kenyon, S. J. 2003, ApJ, 583, 334
Hartmann, L. 2001, AJ, 121, 1030
Herczeg, C. J., & Hillenbrand L. A. 2008, ApJ 681, 594
Herczeg, C. J., et al. 2009, in preparation
Hughes, M. A., Wilner, D. J., Qi, C., & Hogerheijde, M. R. 2008, ApJ
678, 1119
Isella, A., Carpenter, J. M., & Sargent, A. F. 2009, ApJ 701, 260
Itoh, Y., et al. 2009, ApJ 620, 984
Jensen, E. L. N., Dhital, S., Stassun, K. G., Patience, J., Herbst, W.,
Walter, F. M., Simon, M., & Basri, G. 2007, AJ, 134, 241
Jensen, E. L. N., Koerner, D. W., & Mathieu, R.D. 1996, AJ, 111,
2431
Kenyon, S. J., & Hartmann L. 1995, ApJS 101, 117
Kessler-Silacci, J.; Augereau, J. C.; Dullemond, C. P., Geers, V.,
Lahuis, F., et al. 2006, ApJ 639, 275
Kitamura, Y., Momose, M., Yokogawa, S., et al. 2002, ApJ 581, 357
Klahr, H. H., & Henning, T. 1997, Icarus, 128, 213
Klein, R., Apai, D., Pascucci, I., Henning, Th., & Waters, L. B. F. M.
2003, ApJ 593, 57
Koerner, D. W., Chandler, C. J., & Sargent, A. I. 1995, ApJ 452, 69
Kwon, W., Looney, L. W., Mundy, L. G., Chiang H., & Kemball, A.
2009, ApJ 696, 841
Leggett, S. K. 1992, ApJS 82, 351
Leinert, C., Zinnecker, H., Weitzel, N., Christou, J., Ridgway, S. T.,
Jameson, R., Haas, M., & Lenzen, R. 1993, A&A 278, 129
Lommen, D., Maddison, S. T., Wright, C. M., van Dishoeck, E. F.,
Wilner, D. J., & Bourke, T. L. 2009, A&A 495, 869
Lommen, D., Wright, C. M., Maddison, S. T., Jrgensen, J. K., Bourke,
T. L., van Dishoeck, E. F., Hughes, A., Wilner, D. J., Burton, M.,
& van Langevelde, H. J. 2007, A&A 462, 211
Luhman, K. L. 1999, ApJ 525, 466
Malfait, K., Waelkens, C., Waters, L. B. F. M., Vandenbussche, B.,
Huygen, E., & de Graauw, M. S. 1998, A&A 332, 25
Mannings, V., & Emerson, J. P. 1994, MNRAS 267, 361
Mathieu, R.D. 1994, ARA&A 32, 465
Mathis, J. S., Rumpl, W., & Nordsieck, K. H. 1977, ApJ 217, 425
Meyer, M. R., Calvet, N., & Hillenbrand, L. A. 1997, AJ 114, 288
Miyake, K., & Nakagawa, Y. 1993, Icarus 106, 20
Mundy, L. G., et al. 1996, ApJ 464, L169
Natta, A., Testi, L., Neri, R., Shepherd, D. S., & Wilner, D. J. 2004,
A&A 416, 179
Ohashi, N., Hayashi, M., Kawabe, R., & Ishiguro, M. 1996, ApJ 466,
317
|
1904.12790 | 1 | 1904 | 2019-04-29T16:07:27 | Astro 2020: Astromineralogy of interstellar dust with X-ray spectroscopy | [
"astro-ph.EP",
"astro-ph.HE"
] | X-ray absorption fine structure (XAFS) in the 0.2-2 keV band is a crucial component in multi-wavelength studies of dust mineralogy, size, and shape -- parameters that are necessary for interpreting astronomical observations and building physical models across all fields, from cosmology to exoplanets. Despite its importance, many fundamental questions about dust remain open. What is the origin of the dust that suffuses the interstellar medium (ISM)? Where is the missing interstellar oxygen? How does iron, predominantly produced by Type Ia supernovae, become incorporated into dust? What is the main form of carbon in the ISM, and how does it differ from carbon in stellar winds? The next generation of X-ray observatories, employing microcalorimeter technology and $R \equiv \lambda/\Delta \lambda \geq 3000$ gratings, will provide pivotal insights for these questions by measuring XAFS in absorption and scattering. However, lab measurements of mineralogical candidates for astrophysical dust, with R > 1000, are needed to fully take advantage of the coming observations. | astro-ph.EP | astro-ph | Astro2020 Science White Paper
Astromineralogy of interstellar dust with X-ray
spectroscopy
(cid:3) Planetary Systems (cid:52) Star and Planet Formation
Thematic Areas:
(cid:3) Formation and Evolution of Compact Objects
(cid:3) Stars and Stellar Evolution (cid:3) Resolved Stellar Populations and their Environments
(cid:3) Multi-Messenger Astronomy and Astrophysics
(cid:3) Galaxy Evolution
(cid:3) Cosmology and Fundamental Physics
Principal Author:
Name: L´ıa Corrales
Institution: University of Michigan
Email: [email protected]
Phone: (734) 763-8915
Co-authors: Lynne Valencic (Johns Hopkins University); Elisa Costantini (SRON); Javier Garc´ıa
(Caltech); Efrain Gatuzz (MPA); Tim Kallman (GSFC); Julia Lee (Harvard); Norbert Schulz (MIT);
Sascha Zeegers (ASIAA); Claude Canizares (MIT); Bruce Draine (Princeton); Sebastian Heinz
(University of Wisconsin); Edmund Hodges-Kluck (GSFC); Edward B. Jenkins (Princeton); Frits
Paerels (Columbia); Randall K. Smith (SAO); Tea Temim (STScI); Jorn Wilms (University of
Erlangen-Nuremberg); Daniel Wolf Savin (Columbia)
Abstract (optional):
X-ray absorption fine structure (XAFS) in the 0.2 -- 2 keV band is a crucial component in
multi-wavelength studies of dust mineralogy, size, and shape -- parameters that are necessary
for interpreting astronomical observations and building physical models across all fields, from
cosmology to exoplanets. Despite its importance, many fundamental questions about dust remain
open. What is the origin of the dust that suffuses the interstellar medium (ISM)? Where is the
missing interstellar oxygen? How does iron, predominantly produced by Type Ia supernovae,
become incorporated into dust? What is the main form of carbon in the ISM, and how does
it differ from carbon in stellar winds? The next generation of X-ray observatories, employing
microcalorimeter technology and R ≡ λ/∆λ ≥ 3000 gratings, will provide pivotal insights for
these questions by measuring XAFS in absorption and scattering. However, lab measurements of
mineralogical candidates for astrophysical dust, with R > 1000, are needed to fully take advantage
of the coming observations.
9
1
0
2
r
p
A
9
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
9
7
2
1
.
4
0
9
1
:
v
i
X
r
a
1
The effects of cosmic dust can be seen in virtually every field of astrophysics. The cycle of
baryons, the growth of molecules, the physics of cosmic gas, the formation of planets, and the origin
of life all rely upon the formation of solid phase materials in space. Cosmic dust has even been
found in extraordinarily unexpected places: from the hot halo gas surrounding galaxies (Engelbracht
et al., 2006; M´enard et al., 2010) to the filaments of molecular gas in galaxy clusters (Russell et al.,
2014; Russell et al., 2017).
Despite the importance, ubiquity, and decades of study on astrophysical dust, shockingly basic
questions remain. From the perspective of timescales, even the presence of dust in the diffuse
interstellar medium (ISM) remains a mystery. All dust grains injected by AGB stars into the
surrounding ISM will eventually be processed by the destructive force of a supernova shock (Dwek
et al., 2008; Jones et al., 2013; Raymond et al., 2013). However, the majority of refractory elements
that become incorporated into interstellar dust -- Mg, Si, and Fe -- are produced in supernovae
explosions (Dwek, 2016). The dust that formed rapidly in the ejecta may be destroyed later by
the reverse shock, which takes tens of thousands of years to fully heat the remnant (Gall et al.,
2011; Raymond et al., 2013; Gall et al., 2014). The debate continues whether supernovae -- on
average -- produce or destroy dust. The answer will determine what fraction of interstellar solid
matter is formed in the low density environment of the diffuse ISM, through deposition on existing
grain surfaces (Draine, 2009). Thus, the origin and fate of dust in the Universe remains a
fundamental question in astrophysics, which can be answered by measuring the composition,
size, and structure of interstellar dust.
High resolution X-ray spectroscopy provides abundances for all ions, from neutral to H-like
species, in gaseous and solid form, over a broad range of ISM densities (NH ∼ 1020 -- 1024 cm−2).
The signature of solid phase minerals and glasses are imprinted in the shape of the absorption
features in high resolution X-ray spectra (Lee & Ravel, 2005; Lee et al., 2009). Studying these
X-ray absorption fine structure (XAFS) features yield the constituent minerals, crystalline content
(versus amorphous), size, and shape of interstellar dust grains. X-ray spectroscopy can provide a
complete inventory of both the gas and solid phase of the ISM, while directly identifying the
mineral building-blocks of cosmic dust.
1 Open questions in astromineralogy
Astromineralogy rests upon two major pillars of observational astronomy. First, direct gas phase
abundances found via UV/optical absorption line studies, compared to a total abundance table,
determine the fraction in the solid phase (Jenkins, 2009). These studies are limited to low opacity
sight lines as UV/optical light cannot penetrate the dense ISM. Second, IR and radio studies detect
polycyclic aromatic hydrocarbons (PAHs), silicates, and ices in the dense ISM. These species are
studied using rotational and vibrational transitions, which require complex models. Determining the
abundances or grain structures based on these features requires making strong assumptions about
their physical properties (Draine, 2003).
High throughput, high resolution X-ray spectroscopy can reveal both the structure and composi-
tion of interstellar dust. Structures in X-ray spectra imposed by XAFS provide a direct measure
of the spacing between atoms in a crystalline lattice (Lee & Ravel, 2005), while lab measure-
ments of astrophysically relevant materials provide templates for measuring dust components (Lee
et al., 2009; Zeegers et al., 2017; Rogantini et al., 2018). To fully interpret near edge X-ray
photoabsorption features, optical constants derived from lab absorption measurements must
2
be incorporated into models that calculate the grain size- and shape-dependent scattering
properties (Hoffman & Draine, 2016; Corrales et al., 2016; Zeegers et al., 2017).
Progress in studying the cold phases of the ISM through spectroscopic capabilities of Chandra
and XMM-Newton (R ≡ λ/∆λ ∼ 500 − 1000 with effective areas Ae ∼ 100 cm2), made over the
last two decades, has yielded results that are inconsistent with the standard picture of ISM dust
as studied at longer wavelengths. We review open problems and X-ray findings that require new
observations, lab measurements, and theoretical calculations to solve.
Figure 1: Simulated spectra showing the effects of absorption due to dust. Left: 4U 1746-37
(NH = 5 × 1021 cm2) at the oxygen K edge as observed with a telescope with R = 5000 and
Ae = 4000 cm2. If water ice on grains near the diffuse ISM/cloud interface accounts for the
"missing oxygen", this mission will detect it. Right: Cyg X-2 (NH = 1022 cm2) at the Fe L edge,
observed with R = 3000 and Ae = 500 cm2. It will easily distinguish between iron in various
species of magnetic inclusions or silicates, important for grain polarization at longer wavelengths.
1.1 Solving the Oxygen problem (0.5 keV)
Oxygen is the most abundant interstellar metal and is a key ingredient of ices and silicate minerals.
Studies of the gas-phase ISM suggest that ∼30% of oxygen is locked up in solids, a fraction that
increases with decreasing ISM temperatures. In the coldest regions, more oxygen is "missing" from
the gas phase than can be explained by silicate minerals alone, suggesting the presence of icy grain
mantles (Jenkins, 2009). Poteet et al. (2015) have shown that the H2O ice would have to be in very
thick mantles to have escaped detection by IR spectroscopy. Oxygen is thereby an ideal target for
X-ray absorption studies on sightlines where there is missing oxygen. Studies will need to take
into account the effects of large grain sizes on the X-ray extinction cross sections, but these can be
modeled.
Early measurements of oxygen X-ray absorption argued for the appearance of additional K shell
absorption edges (Paerels et al., 2001; Takei et al., 2002) and residual features arising from the
solid phase (de Vries & Costantini, 2009). Later studies demonstrated that many of these features
could be explained by absorption from gas-phase OII and OIII, and by including the effects of
Auger decay (Juett et al., 2004; Garc´ıa et al., 2011; Gorczyca et al., 2013; Gatuzz et al., 2015). A
recent study by Joachimi et al. (2016) also failed to find statistically relevant CO features in X-ray
absorption from a majority of X-ray binary sight lines. Resolving this disagreement between
3
X-ray observations and the canon of ISM dust studies at longer wavelengths is crucial for
deciphering the mystery of "missing oxygen."
The currently available cross-sections with
XAFS for OI K-shell absorption are incorpo-
rated in the amol model in SPEX (Kaastra
et al., 1996). These cross-sections are avail-
able at a variety of resolutions, from R ≤ 1000
(O2, CO, CO2, and heavy metal oxides, Bar-
rus et al., 1979; van Aken et al., 1998) to
R ≥ 5000 (H2O, Hiraya et al., 2001; Par-
ent et al., 2002). Figure 1 (left) shows the
O K-edge from a flagship mission that would
be able to measure the abundance of various
oxygen bearing species along highly absorbed
sight lines in 200 ks.
Figure 2: Example 50 ks astrosilicate O K (Draine,
2003) measurement obtained from the ratio of
a point source (Fps) and halo spectrum (Fh) us-
X-ray missions using microcalorimeter tech-
ing Ae = 500 cm2, R = 1000 (black) or a mi-
nology, creating the X-ray observational equiv-
crocalorimeter (R ≈ 250, Ae = 104 cm2, in red).
alent of an integral field spectrograph, will en-
able high resolution spectroscopy of dust scattering halos, which will exhibit resonant structure
around the photoelectric edge (Figure 2, see also white paper by Valencic et al., 2019). However,
these features cannot be fully interpreted until high resolution lab measurements of XAFS from ices
(containing CO, CO2, CH3OH, and so on) at R ≥ 1000, and the corresponding optical constants,
are obtained.
1.2 How is iron incorporated into the solid phase? (0.7 keV)
The mystery of how iron is incorporated into dust -- its dominant form -- has vexed astronomers for
decades. Type Ia supernovae produce the majority of interstellar iron but do not appear to produce
dust. Therefore, iron dust must form rapidly in the ISM (Dwek, 2016). Olivine is a prime candidate
material for interstellar silicates, and can exhibit a range of Mg to Fe ratios: fayalite contains iron
only (Fe2SiO4) while forsterite contains magnesium only (Mg2SiO4). Learning whether interstellar
iron is bound mostly in silicates, oxides, or magnetic nanoparticle inclusions will uncover crucial
dust formation mechanisms (e.g., simulations by Zhukovska et al., 2016, 2018).
This question has taken on even greater importance recently, as attempts to detect the cosmic
microwave background polarization have shown how little we know about the polarized dusty
Galactic foreground (Barkats et al., 2014; Remazeilles et al., 2016). The polarized emission from
foreground dust can change quickly in the microwave regime and is highly dependent on grain size
and ferric composition, whether iron is primarily in magnetic inclusions (e.g., Fe2O3, Fe3O4) or
silicates (Draine & Fraisse, 2009; Draine & Hensley, 2013; Hoang & Lazarian, 2016).
A MIDEX-class detector with spectral resolution R = 3000 and Ae = 500 cm2 will be able
to distinguish between different forms of iron, as seen in the Fe L edge region simulated with the
amol model from SPEX (Figure 1, right). The current lab measurements of the Fe L region are
available with R ≈ 3500 (Lee et al., 2009), but many commonly used X-ray absorption models
(tbvarabs, ISMabs) incorporate the Fe L edge of pure solid iron at R ∼ 500 (Kortright & Kim,
2000). New lab measurements of Fe K shell (7.1 keV) absorption from astrosilicate materials are
available with R = 14000 (Rogantini et al., 2018), important when NH ≥ 1023 cm−2.
4
Figure 3: Left: Silicate cross-sections for various mineral compositions, using optical constants
from Zeegers et al. (2017), publicly available with 1 eV resolution. Right: Simulated spectra for
GX 340+0, a highly absorbed LMXB with strong Si K shell features. XAFS from a crystalline
olivine (blue) and amorphous pyroxene (red), with microcalorimeter (R ∼ 1000) and R = 3000
resolution, are overlaid on the Chandra spectrum (grey). An exposure time of 50 ks with Ae =
1000 cm2 was used.
1.3 What is the composition and crystalline fraction of silicates? (1.8 keV)
Silicate dust produced by some red giants contain signs of crystallinity (Molster et al., 2002; Guha
Niyogi et al., 2011), yet infrared Si-O spectra from the diffuse ISM primarily indicate amorphous
silicon (Li & Draine, 2001; Kemper et al., 2004), providing strong evidence for significant dust
processing in the ISM.
Lab measurements of common silicate materials have been performed at R ∼ 4000 − 7000
(Zeegers et al., 2017). They show multiple absorption resonances associated with the crystalline
spacing of Si (Figure 3, left). Some of these features are spread out or lost when the crystalline
structure is destroyed, as seen in the features for amorphous pyroxene. The depth of the near-edge
absorption peak also varies depending on the mineral. Using a cross-section that includes the effects
of dust scattering, Zeegers et al. (2017) showed that the spectrum of low mass X-ray binary GX 5-1
contained features consistent with significantly more crystalline olivine than amorphous pyroxene,
with grains as large as 0.5 µm. This signature of crystalline olivine contradicts the IR spectroscopy
paradigm of amorphous silicates dominating the diffuse ISM. Figure 3 (right) shows how high
throughput, high resolution X-ray spectroscopy can better differentiate between the two.
1.4 Mysteries of carbon dust growth and processing (0.3 keV)
Little work has been done on neutral C K shell absorption due to the fact that currently operating
X-ray telescopes have an Ae that drops quickly for E < 0.5 keV. This is unfortunate, as under-
standing the abundance and form of carbonaceous dust is currently one of the most active fields
in astromineralogy. Small graphite and PAHs alone are considered responsible for the 2175 A
absorption bump and emission features ranging from 3.4 − 20 µm (Draine & Lee, 1984; Draine &
Li, 2007). However, models that utilize compound mixtures of silicates, amorphous carbon, and
ices can also describe extinction and infrared emission features from the UV to sub-mm (Zubko
et al., 2004; Jones et al., 2013). Additionally, nanodiamonds are abundant in presolar grains found
in meteorites, and IR spectra of circumstellar material around some Herbig Ae/Be stars also show
5
signatures of nanodiamonds (e.g., Jones et al., 2004; Goto et al., 2009).
These hypotheses can be tested by X-ray observatories with Ae ≥ 500 cm2 and R ≥ 4000, as
demonstrated for the case of methane, benzene, and nanodiamond material (Bilalbegovi´c et al.,
2018). XAFS can also probe the process of hydrogenation of large organic molecules (Reitsma
et al., 2014). Curently, these types of experiments are only feasible with X-ray gratings.
1.5 What is the true abundance of metals in the local ISM?
Gas can be distinguished from solid phases only with high resolution X-ray spectra, providing
a check on absolute abundances in the ISM -- a crucial assumption in measuring depletion. Fur-
thermore, ISM absorption in the 0.3 -- 10 keV band is strongly influenced by the adopted mix of
interstellar metals, affecting X-ray continuum models on the 30% level (Wilms et al., 2000). A para-
metric study of silicate K shell absorption in the spectra of low mass X-ray binaries demonstrated
that the absorption edges are deeper than predicted by current X-ray ISM models, suggesting that Si
abundances need to be increased (Schulz et al., 2016). Balancing the dust and gas budget for ISM
metals is fundamental for testing our understanding of cosmic star formation history and finding
missing baryons, a key science charge for current and future X-ray observatories (Bregman, 2007).
2 The current state and future prospects
Only two X-ray observatories are currently capable of high resolution X-ray spectroscopy: Chandra,
which excels in spectroscopic resolution but whose soft energy effective area has diminished
significantly over time, and XMM-Newton, which does not have the necessary spectral resolution
to distinguish solidly between grain types. Further, they do not have sufficient Ae to allow selection
of sightlines based on dust characteristics; rather, typically only the brightest, most absorbed X-ray
sources are selected for dust studies. The next generation of X-ray telescopes are capable of
Ae ∼ 500 -- 2000 cm2, offering a factor of 10 -- 100 improvement in signal-to-noise and opening a
larger range of sight lines for scientific discovery.
Planned X-ray observatories like the X-ray Imaging and Spectroscopy Mission (XRISM) and
Athena will employ microcalorimeters to measure an X-ray spectrum from every pixel with 2 -- 4 eV
resolution. Such instruments are ideal for measuring spectra from dust scattering halos, which will
also exhibit fine structure that can be used to evaluate dust mineralogy and grain sizes. This effect is
described in more detail in a separate white paper by Valencic et al. (2019).
Unfortunately, at the soft end of the spectrum (< 2 keV), where the majority of XAFS from
interstellar metals appear, microcalorimeters can only achieve R ∼ 250 at the O K edge (R ∼ 900
for Si K). In addition, their effective areas diminish rapidly in the 0.3 -- 0.5 keV range where O
and C absorption signatures arise. Recently developed critical-angle transmission (CAT) gratings
technology can achieve R ∼ 2500 − 104 in the 0.2 -- 1.4 keV bandpass (Heilmann et al., 2017). Such
instruments are ideal for astromineralogy in our own Galaxy, but would also enable us to probe
cold-phase oxygen in quasar absorption line systems out to z = 1 (carbon out to z = 0.2). While
X-ray microcalorimeters will do their part to reveal the mineralogy of silicate dust, the state
of the two most abundant yet mysterious interstellar metals -- carbon and oxygen -- can more
readily be studied with the latest advances in X-ray gratings spectroscopy. Furthermore, high
resolution lab measurements of X-ray absorption from fundamental mineralogical building
blocks, such as ices and organic molecules, are necessary to interpret the observations that
will be made in the next 20 years.
6
References
Barkats, D., Aikin, R., Bischoff, C., et al. 2014,
ApJ, 783, 67
Barrus, D. M., Blake, R. L., Burek, A. J., Cham-
bers, K. C., & Pregenzer, A. L. 1979, PhRvA,
20, 1045
Bilalbegovi´c, G., Maksimovi´c, A., & Valencic,
L. A. 2018, MNRAS, 476, 5358
Bregman, J. N. 2007, ARA&A, 45, 221
Corrales, L. R., Garc´ıa, J., Wilms, J., & Baganoff,
F. 2016, MNRAS, 458, 1345
Series, Vol. 10399, Society of Photo-Optical
Instrumentation Engineers (SPIE) Conference
Series, 1039914
Hiraya, A., Nobusada, K., Simon, M., et al. 2001,
PhRvA, 63, 042705
Hoang, T., & Lazarian, A. 2016, ApJ, 831, 159
Hoffman, J., & Draine, B. T. 2016, ApJ, 817, 139
Jenkins, E. B. 2009, ApJ, 700, 1299
Joachimi, K., Gatuzz, E., Garc´ıa, J. A., & Kall-
man, T. R. 2016, MNRAS, 461, 352
de Vries, C. P., & Costantini, E. 2009, A&A, 497,
Jones, A. P., d'Hendecourt, L. B., Sheu, S.-Y.,
393
Draine, B. T. 2003, ApJ, 598, 1026
Draine, B. T. 2009, in Astronomical Society
of the Pacific Conference Series, Vol. 414,
Cosmic Dust - Near and Far, ed. T. Henning,
E. Grun, & J. Steinacker, 453
Draine, B. T., & Fraisse, A. A. 2009, ApJ, 696, 1
Draine, B. T., & Hensley, B. 2013, ApJ, 765, 159
Draine, B. T., & Lee, H. M. 1984, ApJ, 285, 89
Draine, B. T., & Li, A. 2007, ApJ, 657, 810
Dwek, E. 2016, ApJ, 825, 136
Dwek, E., Arendt, R. G., Bouchet, P., et al. 2008,
ApJ, 676, 1029
et al. 2004, A&A, 416, 235
Jones, A. P., Fanciullo, L., Kohler, M., et al.
2013, A&A, 558, A62
Juett, A. M., Schulz, N. S., & Chakrabarty, D.
2004, ApJ, 612, 308
Kaastra, J. S., Mewe, R., & Nieuwenhuijzen, H.
1996, in UV and X-ray Spectroscopy of Astro-
physical and Laboratory Plasmas, ed. K. Ya-
mashita & T. Watanabe, 411 -- 414
Kemper, F., Vriend, W. J., & Tielens, A. G. G. M.
2004, ApJ, 609, 826
Kortright, J. B., & Kim, S.-K. 2000, Phys. Rev. B,
62, 12216
Engelbracht, C. W., Kundurthy, P., Gordon,
K. D., et al. 2006, ApJL, 642, L127
Lee, J. C., & Ravel, B. 2005, ApJ, 622, 970
Lee, J. C., Xiang, J., Ravel, B., Kortright, J., &
Gall, C., Hjorth, J., & Andersen, A. C. 2011,
Flanagan, K. 2009, ApJ, 702, 970
A&A Rv, 19, 43
Gall, C., Hjorth, J., Watson, D., et al. 2014, Na-
ture, 511, 326
Garc´ıa, J., Ram´ırez, J. M., Kallman, T. R., et al.
2011, ApJL, 731, L15
Gatuzz, E., Garc´ıa, J., Kallman, T. R., Mendoza,
Li, A., & Draine, B. T. 2001, ApJL, 550, L213
M´enard, B., Scranton, R., Fukugita, M., &
Richards, G. 2010, MNRAS, 405, 1025
Molster, F. J., Waters, L. B. F. M., & Tielens,
A. G. G. M. 2002, A&A, 382, 222
Paerels, F., Brinkman, A. C., van der Meer,
C., & Gorczyca, T. W. 2015, ApJ, 800, 29
R. L. J., et al. 2001, ApJ, 546, 338
Gorczyca, T. W., Bautista, M. A., Hasoglu, M. F.,
Parent, P., Laffon, C., Mangeney, C., Bournel, F.,
et al. 2013, ApJ, 779, 78
& Tronc, M. 2002, JChPh, 117, 10842
Goto, M., Henning, T., Kouchi, A., et al. 2009,
Poteet, C. A., Whittet, D. C. B., & Draine, B. T.
ApJ, 693, 610
2015, ApJ, 801, 110
Guha Niyogi, S., Speck, A. K., & Onaka, T. 2011,
Raymond, J. C., Ghavamian, P., Williams, B. J.,
ApJ, 733, 93
et al. 2013, ApJ, 778, 161
Heilmann, R. K., Bruccoleri, A. R., Song, J.,
et al. 2017, in Society of Photo-Optical In-
strumentation Engineers (SPIE) Conference
Reitsma, G., Boschman, L., Deuzeman, M. J.,
et al. 2014, Physical Review Letters, 113,
053002
7
Remazeilles, M., Dickinson, C., Eriksen,
H. K. K., & Wehus, I. K. 2016, MNRAS, 458,
2032
Rogantini, D., Costantini, E., Zeegers, S. T., et al.
van Aken, P. A., Liebscher, B., & Styrsa, V. J.
1998, Physics and Chemistry of Minerals, 25,
494
Wilms, J., Allen, A., & McCray, R. 2000, ApJ,
2018, A&A, 609, A22
542, 914
Russell, H. R., McNamara, B. R., Edge, A. C.,
Zeegers, S. T., Costantini, E., de Vries, C. P.,
et al. 2014, ApJ, 784, 78
Russell, H. R., McDonald, M., McNamara, B. R.,
et al. 2017, The Astrophysical Journal, 836,
130
Schulz, N. S., Corrales, L., & Canizares, C. R.
et al. 2017, A&A, 599, A117
Zhukovska, S., Dobbs, C., Jenkins, E. B., &
Klessen, R. S. 2016, ApJ, 831, 147
Zhukovska, S., Henning, T., & Dobbs, C. 2018,
ApJ, 857, 94
2016, ApJ, 827, 49
Zubko, V., Dwek, E., & Arendt, R. G. 2004,
Takei, Y., Fujimoto, R., Mitsuda, K., & Onaka,
ApJS, 152, 211
T. 2002, ApJ, 581, 307
8
|
0903.4700 | 1 | 0903 | 2009-03-26T20:59:37 | Planet-planet scattering leads to tightly packed planetary systems | [
"astro-ph.EP"
] | The known extrasolar multiple-planet systems share a surprising dynamical attribute: they cluster just beyond the Hill stability boundary. Here we show that the planet-planet scattering model, which naturally explains the observed exoplanet eccentricity distribution, can reproduce the observed distribution of dynamical configurations. We calculated how each of our scattered systems would appear over an appropriate range of viewing geometries; as Hill stability is weakly dependent on the masses, the mass-inclination degeneracy does not significantly affect our results. We consider a wide range of initial planetary mass distributions and find that some are poor fits to the observed systems. In fact, many of our scattering experiments overproduce systems very close to the stability boundary. The distribution of dynamical configurations of two-planet systems actually may provide better discrimination between scattering models than the distribution of eccentricity. Our results imply that, at least in their inner regions which are weakly affected by gas or planetesimal disks, planetary systems should be "packed", with no large gaps between planets. | astro-ph.EP | astro-ph |
Draft version August 10, 2018
Preprint typeset using LATEX style emulateapj v. 12/14/05
PLANET-PLANET SCATTERING LEADS TO TIGHTLY PACKED PLANETARY SYSTEMS
Sean N. Raymond1,2, Rory Barnes2,3, Dimitri Veras4, Philip J. Armitage5, Noel Gorelick6 & Richard
Greenberg7
(Received; Accepted)
Draft version August 10, 2018
ABSTRACT
The known extrasolar multiple-planet systems share a surprising dynamical attribute: they cluster
just beyond the Hill stability boundary. Here we show that the planet-planet scattering model,
which naturally explains the observed exoplanet eccentricity distribution, can reproduce the observed
distribution of dynamical configurations. We calculated how each of our scattered systems would
appear over an appropriate range of viewing geometries; as Hill stability is weakly dependent on
the masses, the mass-inclination degeneracy does not significantly affect our results. We consider a
wide range of initial planetary mass distributions and find that some are poor fits to the observed
systems. In fact, many of our scattering experiments overproduce systems very close to the stability
boundary. The distribution of dynamical configurations of two-planet systems actually may provide
better discrimination between scattering models than the distribution of eccentricity. Our results
imply that, at least in their inner regions which are weakly affected by gas or planetesimal disks,
planetary systems should be "packed", with no large gaps between planets.
Subject headings: planetary systems: formation -- methods: n-body simulations
1. INTRODUCTION
The observed eccentricities of extra-solar planets can
be readily explained by a simple model that assumes
that virtually all planetary systems undergo dynami-
cal instabilities (Ford et al. 2003; Adams & Laughlin
2003; Chatterjee et al. 2008; Juric & Tremaine 2008;
Ford & Rasio 2008).8
In the context of this model,
planetary systems are expected to form in marginally
stable configurations, meaning that they are stable for
at least the timescale of rapid gas accretion of ∼ 105
years (Pollack et al. 1996) but ultimately unstable, prob-
ably on a timescale comparable to the gaseous disk's
lifetime of ∼ 106 years (Haisch et al. 2001). This in-
stability timescale implies an initial separation between
planets of perhaps 4-5 mutual Hill radii RH,M , where
RH,M = 0.5 (a1 + a2)[(M1 + M2)/3M⋆]1/3; a1 and a2 are
the orbital distances, M1 and M2 are the masses of two
adjacent planets, and M⋆ is the stellar mass (Chambers
et al. 1996; Marzari & Weidenschilling 2002; Chatterjee
et al. 2008).9 After a delay of 105 − 106 years, a typi-
1 Center
for
389 UCB, University
[email protected]
Astrophysics
and
Astronomy,
of Colorado, Boulder CO 80309;
Space
2 Virtual Planetary Laboratory
3 Department of Astronomy, University of Washington, Seattle,
WA 98195
4 Astronomy Department, University of Florida, Gainesville,
FL 32111
5 JILA, University of Colorado, Boulder CO 80309
6 Google, Inc., 1600 Amphitheatre Parkway, Mountain View,
CA 94043
7 Lunar and Planetary Laboratory, University of Arizona,
Tucson, AZ
8 Several other models to explain the extra-solar eccentricity
distribution exist; see Ford & Rasio (2008) for a summary.
9 For Jupiter-mass planets, separations of ∼ 4 −5RH,M are close
to the 3:2 and 2:1 mean motion resonances. Thus, an alternate
argument in favor of planets forming with such spacings invokes
resonant capture (Snellgrove et al. 2001) followed by turbulent re-
moval from resonance (Adams et al. 2008) during the gaseous disk
phase.
cal system of three or more planets with separations of
4 − 5RH,M becomes unstable, leading to close encoun-
ters between two planets, strong dynamical scattering,
and eventual destruction of one or two planets by either
collision with another planet, collision with the star, or,
most probably, hyperbolic ejection from the system (Ra-
sio & Ford 1996; Weidenschilling & Marzari 1996; Lin &
Ida 1997; Papaloizou & Terquem 2001). It is the plan-
ets that survive the dynamical instability that provide a
match to the observed extra-solar eccentricities.
Additional dynamical
information can be obtained
from the known extra-solar multiple planet systems. In
particular, the stability in two-planet systems can be
guaranteed for planets with particular masses and orbital
configurations. The edge of stability can be quantified in
terms of the proximity to the analytically-derived Hill
stability limit using the dimensionless quantity β/βcrit
(the stability boundary is located at β/βcrit = 1; see
Section 3). Dynamical analyses have shown that the
known multiple-planet systems are clustered just beyond
the edge of stability (i.e., at β/βcrit & 1; Barnes & Quinn
2004; Barnes & Greenberg 2006, 2007).
In this paper we study the stability of the surviving
planets in several thousand 3-planet systems that have
undergone planet-planet scattering leading to the loss of
one planet. We find that in the aftermath of dynamical
instabilities, the surviving planets cluster just beyond the
stability boundary, providing a good match to the ob-
served values. This provides support for planet-planet
scattering as an active process in extra-solar planetary
systems. This result also has consequences for the pack-
ing of planetary systems and the "Packed Planetary Sys-
tems" hypothesis (Barnes & Raymond 2004; Raymond
& Barnes 2005; Raymond et al. 2006; Barnes et al. 2008).
The paper proceeds as follows: we describe our scatter-
ing simulations (§2), summarize Hill stability theory and
define β/βcrit (§3), present our results (§4) and discuss
the consequences (§5).
2
Raymond et al.
2. SCATTERING SIMULATIONS
Our scattering simulations are drawn from the same
sample as in Raymond et al. (2008a). Each simulation
started with three planets randomly separated by 4-5
mutual Hill radii. The three planets were placed such
that the outermost planet was located two (linear) Hill
radii RH interior to 10 AU (RH = a[M/3M⋆]1/3). We
performed ten sets of simulations, varying the planetary
mass distribution. For our two largest sets (1000 sim-
ulations each) we randomly selected planet masses ac-
cording to the observed distribution of exoplanet masses:
dN/dM ∝ M −1.1 (Butler et al. 2006). In the Mixed1 set
we restricted the planet mass Mp to be between a Sat-
urn mass MSat and three Jupiter masses MJup. For our
Mixed2 set, the minimum planet mass was decreased to
10 M⊕. We also performed four Meq sets (500 simula-
tions each) with equal mass planets for Mp = 30 M⊕,
MSat, MJup, and 3MJup. Finally, the Mgrad sets (250
simulations each) contained radial gradients in Mp. For
the JSN set, in order of increasing orbital distance, Mp =
MJup, MSat, and 30 M⊕. For the NSJ set, these masses
were reversed, i.e., the MJup planet was the most dis-
tant. The 3JJS and SJ3J sets had, in increasing radial
distance, Mp = 3MJup, MJup and MSat, and Mp = MSat,
MJup and 3MJup, respectively.
Planetary orbits were given zero eccentricity and mu-
tual inclinations of less than 1 degree. Each simulation
was integrated for 100 Myr with the hybrid Mercury in-
tegrator (Chambers 1999) using a 20 day timestep. We
required that all simulations conserve energy to better
than dE/E < 10−4, which is needed to accurately test
for stability (Barnes & Quinn 2004). We achieved this
by reducing the timestep to 5 days for simulations with
dE/E > 10−4 and then removing simulations that still
conserved energy poorly. As expected, these systems
were typically unstable on 105 − 106 year timescales. In
addition, about 1/4 of simulations were stable for 100
Myr which shows that we started close to the stability
boundary. For this paper, we restrict our analysis to the
subsample of simulations that 1) were unstable, and 2)
contained two planets on stable orbits after 100 Myr (i.e.,
one and only one planet was destroyed).
3. HILL STABILITY
For the case of two planets with masses M1 and M2
orbiting a star, dynamical stability is guaranteed if:
G2(M1M2 + M⋆M1 + M⋆M2)3 c2h ≥
M1M2(11M1 + 7M2)
M1M2
(M1 + M2)4/3
−2(M⋆ + M1 + M2)
3M⋆(M1 + M2)2
−
1 + 34/3
M 2/3
⋆
,(1)
where c and h represent the total orbital angular mo-
mentum and energy of the system, respectively (Marchal
& Bozis 1982; Gladman 1993; Veras & Armitage 2004;
note that this definition assumes that M1 > M2). We
refer to the left side of Eqn 1 as β and the right side as
βcrit (Barnes & Greenberg 2006). The quantity β/βcrit
therefore measures the proximity of a pair of orbits to
the Hill stability limit of β/βcrit = 1. We note that our
β/βcrit analysis only applies for two-planet non-resonant
systems, because perturbations from additional compan-
ions can shift the stability boundary to values other than
1 (Barnes & Greenberg 2007).
Fig.
1. -- Cumulative distribution of β/βcrit of the well-
characterized extra-solar multi-planet systems (in gray; see Table
1), as compared with our scattering simulations.
When calculating β/βcrit for extra-solar systems, past
research (Barnes & Greenberg 2006, 2007) has assumed
coplanar orbits with masses equal to minimum masses.
Those values of β/βcrit were systematically affected by
the mass-inclination degeneracy, probably resulting in
overestimations. In contrast, our simulations provide the
full three-dimensional orbits, and hence we can calcu-
late the true value of β/βcrit. More importantly, if we
assume that viewing geometries are distributed isotropi-
cally (i.e. edge-on systems are more likely than face-on),
we can determine how β/βcrit would be calculated from
radial velocity data (e.g. assuming coplanar, edge-on or-
bits). For example, if two planets with masses Mb and
Mc have inclinations (relative to their invariable plane)
ib and ic, and the inclination to the line of sight is I, then
the "observed" β/βcrit would use masses Mb sin(ib + I)
and Mc sin(ic + I). In §4 we use this approach to build
a distribution of β/βcrit that is directly comparable to
the actual distribution (and effectively break the mass-
inclination degeneracy).
4. RESULTS
We generated β/βcrit distributions from our simula-
tions following the procedure described above. First, we
"observed" each system from 100 viewing angles, thereby
decreasing the inferred mass of each planet by a factor
of sin(I + ij), where ij refers to each planet's inclination
with respect to a fiducial plane. Second, we assumed the
observed systems to be coplanar in calculating β/βcrit
for each viewing angle using Eqn 1. Finally, we included
the β/βcrit calculated for each viewing angle by assum-
ing the viewing angle I to be isotropically distributed.
Figure 1 compares the cumulative β/βcrit distributions
for the observed two-planet systems with our scatter-
ing simulations. It is important to note that the "true"
β/βcrit distributions, calculated with knowledge of the
simulated systems' real masses and inclinations, are vir-
tually identical to the curves from Fig. 1 (this issue is dis-
cussed further in §5). Table 1 lists the extra-solar systems
in our analysis; we excluded systems with controversial
or poorly-characterized orbits and those that were likely
affected by tidal effects.
In a two planet system with
an inner planet at . 0.1 AU, tides will decrease the in-
ner planet's eccentricity and semimajor axis (Jackson et
Planetary systems included in β/βcrit analysis1
p values from K-S tests of observations vs. scattering
TABLE 1
TABLE 2
simulations
3
System
(pair)
a1, a2
(AU)
e1, e2
HD 202206 b-c2
HD 82943 c-b2
HD 128311 b-c2
HD 73526 b-c2
HD 45364 b-c2
47 UMa b-c
HD 155358 b-c
HD 177830 c-b
HD 60532 b-c2
HD 183263 b-c
HD 108874 b-c2
HD 12661 b-c
HD 11506 c-b
HD 208487 b-c
HD 169830 b-c
HD 168443 b-c
HD 38529 b-c
HD 47186 b-c
0.83,2.55
0.746,1.19
1.099,1.76
0.66,1.05
0.681,0.897
2.11,3.39
0.628,1.224
0.514,1.22
0.77,1.58
1.52,4.25
1.051,2.68
0.83,2.56
0.639,2.43
0.49,1.8
0.81,3.60
0.3,2.91
0.129,3.68
0.05,2.395
0.435,0.267
0.359,0.219
0.25,0.17
0.19,0.14
0.168,0.097
0.049,0.22
0.112,0.176
0.40,0.041
0.278,0.038
0.38,0.253
0.07,0.25
0.35,0.2
0.42,0.22
0.32,0.19
0.31,0.33
0.529,0.212
0.29,0.36
0.038,0.249
M1, M2
(MJ up)
17.4,2.44
2.01,1.75
2.18,3.21
2.9,2.5
0.187,0.658
2.6,0.46
0.89,0.504
0.186,1.43
3.15,7.46
3.69,3.82
1.36,1.018
2.3,1.57
0.82,3.44
0.45,0.46
2.88,4.04
8.02,18.1
0.78,12.7
0.072,0.35
β/βcrit
0.883
0.946
0.968
0.982
0.989
1.025
1.043
1.046
1.054
1.066
1.10
1.12
1.17
1.20
1.28
1.95
2.06
6.13
Case
Mixed1
Mixed2
Meq:3MJup
Meq:MJup
Meq:MSat
Meq:30M⊕
Mgrad:JSN
Mgrad:NSJ
Mgrad:3JJS
Mgrad:SJ3J
All 10 cases
p
0.14
0.13
1.8 × 10−6
9.3 × 10−4
0.12
0.29
4.0 × 10−3
1.5 × 10−4
9.1 × 10−3
4.9 × 10−3
0.14
p (β/βcrit ≤ 1)
p (β/βcrit > 1)
6.2 × 10−3
1.2 × 10−4
2.1 × 10−5
0.10
0.81
0.60
0.12
0.10
1.9 × 10−5
6.2 × 10−4
5.1 × 10−3
0.53
0.07
4.6 × 10−4
0.02
0.43
0.89
2.4 × 10−4
0.02
9.8 × 10−4
2.8 × 10−4
0.58
This contrasts with resonances generated by convergent
migration in gaseous disks, which tend to exhibit low
amplitude libration of more than one resonant argument
(Snellgrove et al. 2001; Lee & Peale 2002).
aSee http://www.astro.washington.edu/users/rory/research/xsp/dynamics/
for an up to date list of β/βcrit values for the known extra-solar
multiple planet systems.
Orbital values were retrieved from
http://exoplanet.eu and http://exoplanets.org.
bThese systems have been claimed to be in mean motion resonances.
al. 2008), thereby increasing the separation between the
two planets and β/βcrit.
Four individual cases -- Mixed1, Mixed2, Meq:MSat,
and Meq:30 M⊕ -- each provide a match to the observed
β/βcrit distribution. Kolmogorov-Smirnov (K-S) tests
show that the probability p that the β/βcrit distributions
from those four cases are drawn from the same distribu-
tion as the observed sample are all 0.1 or larger (Table
2). The distribution calculated by an unweighted com-
bination of all ten cases is also a good match (each case
was given equal weight, regardless of the number of sim-
ulations).
All of our sets of simulation produced a smaller fraction
of systems at β/βcrit < 1 than for the observed systems.
We therefore calculated K-S p values by confining the
distributions to the ranges β/βcrit ≤ 1 and β/βcrit > 1.
All but one case with p ≥ 0.1 also had p (β/βcrit) > 0.1
(Mixed2; see Table 2). However, some cases provide
good matches for β/βcrit ≤ 1) but not for other regions,
notably Meq:MJup, Mgrad:JSN, and Mgrad:NSJ. Systems
with β/βcrit < 1 are unusual because they lie within the
formal Hill stability boundary but are stabilized by spe-
cial orbital configurations. In fact, all five of the known
exoplanet systems with β/βcrit < 1 are thought to lie in
mean motion resonances (Table 1). The scattered sys-
tems with β/βcrit < 1 are stabilized by resonances or
in many cases by low-amplitude, aligned apsidal libra-
tion. Four cases in our sample generated resonant sys-
tems in at least 5% of simulations (Raymond et al. 2008a)
-- Mixed2, Mgrad:JSN, Mgrad:NSJ, and Mgrad:SJ3J --
but only two of these have p (β/βcrit ≤ 1) > 0.1. We
attribute the lack of a correlation between resonances
and β/βcrit < 1 to the relative weakness of these res-
onances. Indeed, resonances caused by scattering tend
to exhibit relatively high-amplitude libration of only one
resonant argument (Raymond et al. 2008a); these reso-
nances have typical β/βcrit values of slightly more than
1 (median β/βcrit = 1.01 − 1.03 for the different cases).
Simulations with radial mass gradients (Mgrad) over-
produced systems very close to the stability boundary,
while cases with equal masses (Meq) produced much
larger β/βcrit values (Fig. 1). A similar effect was seen
in the eccentricity distributions: the Mgrad cases yielded
much smaller eccentricities than the Meq cases (Raymond
et al. 2008a; see also Ford et al. 2003). The Mixed1 and
Mixed2 cases fall between these two regimes. These
trends can be explained by the number of close encoun-
ters nenc that occur in the different cases before a planet
is destroyed. For the Mgrad cases nenc is typically be-
tween 30 and 80, and is larger for less massive sys-
tems (JSN and NSJ). For the Meq cases nenc is vastly
larger, with median values between 100 (3MJup) and 2000
(30 M⊕). The larger number of scattering events in-
creases the eccentricity of surviving planets and also
causes the systems to spread out farther.
In calculating "observed" β/βcrit distributions from
our simulations, we assumed that the viewing angles I
were isotropically distributed. Given that known extra-
solar planet systems are each observed at a fixed I, could
this have introduced a bias in our samples? Figure 2
shows the inferred value of β/βcrit as a function of I
for five Mixed1 systems with varying mutual inclinations
∆i. For ∆i . 35◦, β/βcrit varies only slightly with the
viewing angle, but for large ∆i β/βcrit can change sub-
stantially with I.10 However, β/βcrit varies by more than
10% [20%] over the entire range of possible viewing an-
gles for fewer than 10% [2%] of cases. For all systems,
the edge-on β/βcrit values agree with the true β/βcrit
values (calculated with knowledge of the planets' true
masses and orbits) to better than 10%. There is a small
bias: ∼80% of systems exhibit a shallow negative slope
in β/βcrit vs. I, suggesting that the majority of inferred
β/βcrit values may be overestimated but only by . 1%.
Thus, although I and ∆i are important to keep in mind,
they introduce a negligible error into our analysis.
5. DISCUSSION
The planet-planet scattering model appears to be con-
sistent with the β/βcrit distribution of the observed ex-
10 We have found that it is actually the angular momentum
deficit (Laskar 1997) which controls the magnitude of β/βcrit vari-
ation with I.
4
Raymond et al.
HD 74156 is an example of a packed planetary sys-
tem. Prior to 2008, two planets were known in the
system, at 0.28 and 3.4 AU (Naef et al. 2004), with
β/βcrit = 1.987. Raymond & Barnes (2005) mapped
out a narrow stable zone between the two planets, from
0.9-1.4 AU. The planet HD 74156 d was discovered three
years later by Bean et al. (2008) at 1.01 AU (see also
Barnes et al. 2008) at the peak of the stable zone. We
therefore expect additional planets to exist in systems
with β/βcrit > 1.5 − 2, notably HD 38529 (Raymond
& Barnes 2005) and HD 47186 (Kopparapu et al. 2009).
The probable location of additional planets can be deter-
mined using test planets to map out dynamically stable
regions between known planets (e.g., Menou & Tabach-
nik 2003; Rivera & Haghighipour 2007; Raymond et
al. 2008b).
The β/βcrit distribution of the observed extra-solar
planetary systems may contain information about dif-
ferent dynamical regimes. The region of β/βcrit ≤ 1
is populated entirely by resonant systems and may pro-
vide evidence of planetary system compression, presum-
ably via convergent migration in gaseous protoplane-
tary disks (Snellgrove et al. 2001; Lee & Peale 2002).
The region of 1 ≤ β/βcrit ≤ 1.5 − 2 is consistent with
the scattering regime. Widely-separated systems with
β/βcrit > 1.5 − 2 may have been drawn apart by inter-
actions with planetesimal or gaseous disks (e.g., Gomes
et al. 2004; Moeckel et al. 2008). However, this seems
unlikely given that the known planets lie relatively close
to their stars and that disk effects should be far more
pronounced at large distances.
Given that our simulations started with only three
planets, we could not calculate β/βcrit values in per-
turbed two-planet systems. For example, an interest-
ing comparison with observations would be to measure
β/βcrit for the two easiest-to-detect planets in scattered
three planet systems. This would address the question
of whether to search for additional planets in between
or interior/exterior to the known planets in two planet
systems with large β/βcrit.
We thank Google for access to their machines. S.N.R.
and R.B. acknowledge funding from NASA Astrobiol-
ogy Institutes's Virtual Planetary Laboratory lead team,
supported by NASA under Cooperative Agreement No.
NNH05ZDA001C.
planets ranges from <2 AU to >15 AU.
Fig. 2. -- Inferred values for β/βcrit as a function of observation
angle I for several examples from the Mixed1 set, labeled by the
approximate mutual inclination ∆i between planets. One resonant
it has ∆i = 9◦. I = 90◦ is edge-on and
case is labeled "2:1":
I = 0◦ is face-on.
oplanet systems. The distribution can be reasonably re-
produced by several of our sets of simulations, or even by
an unweighted combination of all ten sets. We therefore
cannot strongly constrain the initial planetary mass dis-
tribution, although we can rule out cases with very poor
fits -- Meq:3MJup, Mgrad:3JJS and Mgrad:SJ3J in partic-
ular -- as the major contributors to the distribution (see
Table 2). We consider the Mixed1 set to be the most re-
alistic because it is drawn from the observed mass distri-
bution (Butler et al. 2006), and it provides a good match
to the observed eccentricity distribution (Raymond et
al. 2008a). In the coming years, we expect many more
systems to be discovered with β/βcrit ≈ 1 − 1.5.
The pileup of scattered systems just beyond the
stability boundary implies that planetary systems are
"packed", meaning that large spaces in between plan-
ets should be rare.11 This provides a theoretical foun-
dation for the "Packed Planetary Systems" hypothesis,
which asserts that if a stable zone exists between two
known planets, then that zone is likely to contain a planet
(Barnes & Raymond 2004; Raymond & Barnes 2005;
Raymond et al. 2006). Given the small β/βcrit values
of scattered systems, there is simply no room to insert
another planet between the two known planets without
causing the system to be unstable.
11 It is important to note that the spacing for planets with
β/βcrit ≈ 1 can be large. Among just the Mixed1 simulations with
1 ≤ β/βcrit ≤ 1.1 the difference in semimajor axis for adjacent
REFERENCES
Adams, F. C., & Laughlin, G. 2003, Icarus, 163, 290
Adams, F. C., Laughlin, G., & Bloch, A. M. 2008, ApJ, 683, 1117
Barnes, R., Go´zdziewski, K., & Raymond, S. N. 2008, ApJ, 680,
L57
Barnes, R., & Greenberg, R. 2007, ApJ, 665, L67
Barnes, R., & Greenberg, R. 2006, ApJ, 647, L163
Barnes, R., & Quinn, T. 2004, ApJ, 611, 494
Barnes, R., & Raymond, S. N. 2004, ApJ, 617, 569
Bean, J. L., McArthur, B. E., Benedict, G. F., & Armstrong, A.
2008, ApJ, 672, 1202
Butler, R. P., et al. 2006, ApJ, 646, 505
Chambers, J. E. 1999, MNRAS, 304, 793
Chambers, J. E., Wetherill, G. W., & Boss, A. P. 1996, Icarus, 119,
261
Chatterjee, S., Ford, E. B., Matsumura, S., & Rasio, F. A. 2008,
ApJ, 686, 580
Ford, E. B., Rasio, F. A., & Yu, K. 2003, Scientific Frontiers in
Research on Extrasolar Planets, 294, 181
Ford, E. B., & Rasio, F. A. 2008, ApJ, 686, 621
Haisch, K. E., Jr., Lada, E. A., & Lada, C. J. 2001, ApJ, 553, L153
Jackson, B., Greenberg, R., & Barnes, R. 2008, ApJ, 678, 1396
Juri´c, M., & Tremaine, S. 2008, ApJ, 686, 603
Kopparapu, R., Raymond, S. N., & Barnes, R. 2009, ApJ, in press,
arXiv:0903.3597
Laskar, J. 1997, A&A, 317, L75
Lin, D. N. C., & Ida, S. 1997, ApJ, 477, 781
Marzari, F., & Weidenschilling, S. J. 2002, Icarus, 156, 570
Menou, K., & Tabachnik, S. 2003, ApJ, 583, 473
Moeckel, N., Raymond, S. N., & Armitage, P. J. 2008, ApJ, 688,
1361
Rasio, F. A., & Ford, E. B. 1996, Science, 274, 954
Raymond, S. N., & Barnes, R. 2005, ApJ, 619, 549
Raymond, S. N., Barnes, R., Armitage, P. J., & Gorelick, N. 2008a,
ApJ, 687, L107
Raymond, S. N., Barnes, R., & Gorelick, N. 2008b, ApJ, 689, 478
Raymond, S. N., Barnes, R., & Kaib, N. A. 2006, ApJ, 644, 1223
Rivera, E., & Haghighipour, N. 2007, MNRAS, 374, 599
Snellgrove, M. D., Papaloizou, J. C. B., & Nelson, R. P. 2001, A&A,
Naef, D., Mayor, M., Beuzit, J. L., Perrier, C., Queloz, D., Silvan,
374, 1092
J. P., & Udry, S. 2004, A&A, 414, 351
Papaloizou, J. C. B., & Terquem, C. 2001, MNRAS, 325, 221
Pollack, J. B., Hubickyj, O., Bodenheimer, P., Lissauer, J. J.,
Podolak, M., & Greenzweig, Y. 1996, Icarus, 124, 62
Veras, D., & Armitage, P. J. 2004, Icarus, 172, 349
Weidenschilling, S. J., & Marzari, F. 1996, Nature, 384, 619
5
|
1801.07333 | 1 | 1801 | 2018-01-16T16:18:37 | Life Beyond the Solar System: Space Weather and Its Impact on Habitable Worlds | [
"astro-ph.EP",
"astro-ph.IM",
"astro-ph.SR"
] | The search of life in the Universe is a fundamental problem of astrobiology and a major priority for NASA. A key area of major progress since the NASA Astrobiology Strategy 2015 (NAS15) has been a shift from the exoplanet discovery phase to a phase of characterization and modeling of the physics and chemistry of exoplanetary atmospheres, and the development of observational strategies for the search for life in the Universe by combining expertise from four NASA science disciplines including heliophysics, astrophysics, planetary science and Earth science. The NASA Nexus for Exoplanetary System Science (NExSS) has provided an efficient environment for such interdisciplinary studies. Solar flares, coronal mass ejections and solar energetic particles produce disturbances in interplanetary space collectively referred to as space weather, which interacts with the Earth upper atmosphere and causes dramatic impact on space and ground-based technological systems. Exoplanets within close in habitable zones around M dwarfs and other active stars are exposed to extreme ionizing radiation fluxes, thus making exoplanetary space weather (ESW) effects a crucial factor of habitability. In this paper, we describe the recent developments and provide recommendations in this interdisciplinary effort with the focus on the impacts of ESW on habitability, and the prospects for future progress in searching for signs of life in the Universe as the outcome of the NExSS workshop held in Nov 29 - Dec 2, 2016, New Orleans, LA. This is one of five Life Beyond the Solar System white papers submitted by NExSS to the National Academy of Sciences in support of the Astrobiology Science Strategy for the Search for Life in the Universe. | astro-ph.EP | astro-ph | Life Beyond the Solar System: Space Weather and Its Impact on Habitable Worlds
Airapetian, V. S.1,2, Danchi, W. C.1, Dong, C. F.3, Rugheimer, S. 4, Mlynczak, M. 5 ,
Stevenson, K. B.6, Henning, W. G.1,7, Grenfell, J. L.8 , Jin, M.9, Glocer, A.1, Gronoff,
G.5,10, Lynch, B.11, Johnstone, C.12, Lüftinger, T.12, Güdel,M.12, Kobayashi, K.13,
Fahrenbach, A. 14, Hallinan, G.15, Stamenkovic, V.16, Cohen, O.17, Kuang, W.1, van der
Holst, B 18, Manchester, C.18, Zank, G.19, Verkhoglyadova, O. 16, Sojka, J. 19, Maehara,
H.20, Notsu, Y.21, Yamashiki, Y.22, France, K.23, Lopez Puertas, M.24, Funke, B.24,
Jackman, C.1, Kay, C 1, Leisawitz, D.1, Alexander, D.25
1 NASA Goddard Space Flight Center, MD 2 American University, DC 3 Princeton
University, NJ, 4 University of St. Andrews, 5 NASA Langley Research Center, VA, 6 Space
Telescope Science Institute, MD, 7 University of Maryland, MD, 8 German Aerospace
Centre, Berlin, 9 UCAR/LMRC, 10 SSAI, VA, 11 SSL/UC Berkley, CA, 12 University of
Vienna, Austria, 13 Yokohama National University, Japan, 14 ELSI/Tokyo Tech, Japan, 15
California Institute of Technology, Pasadena, CA, 16 JPL/California Institute of
Technology, Pasadena, CA, 17 University of Massachusetts, Lowell, MA 18 University of
Michigan, Ann Harbor, MI, 18 University of Alabama, Huntsville, AL, 19 University of
Utah, UT, 20 National Astronomical Observatory of Japan, Japan, 20 Kyoto University,
Japan, 23 University of Colorado at Boulder, CO, 24 Institute de Astrofisica de Andalucía,
Spain, 25Rice University, TX.
Submitted to the National Academy of Sciences in support of the Astrobiology Science
Strategy for the Search for Life in the Universe
NASA's Nexus for Exoplanet System Science (NExSS) is a research coordination network
dedicated to the study of planetary habitability using a system science approach with inputs
from astrophysics, Earth science, planetary science, and heliophysics. Herein, the NExSS
community describes recent progress and future prospects for characterization and
modeling of exoplanetary systems and technology development required to detect and
identify signs of life.
Introduction.
The search of life in the Universe is a fundamental problem of astrobiology and a
major priority for NASA. A key area of major progress since the NASA Astrobiology
Strategy 2015 (NAS15) has been a shift from the exoplanet discovery phase to a phase of
characterization and modeling of the physics and chemistry of exoplanetary atmospheres,
and the development of observational strategies for the search for life in the Universe by
combining expertise from four NASA science disciplines (heliophysics, astrophysics,
planetary science and Earth science, HAPE community). The NASA Nexus for
Exoplanetary System Science (NExSS) has provided an efficient environment for such
interdisciplinary studies.
Solar flares, coronal mass ejections (CMEs) and solar energetic particles (SEPs)
produce disturbances in interplanetary space collectively referred to as space weather,
which interacts with the Earth's upper atmosphere and causes dramatic impact on space-
and ground-based technological systems [1]. Exoplanets within close-in habitable zones
(HZs) around M dwarfs are exposed to extreme ionizing radiation fluxes, thus making
exoplanetary space weather (ESW) effects a crucial factor of habitability [2,3]. In this
paper, we describe the recent developments and provide recommendations in this
interdisciplinary effort with the focus on the impacts of ESW on habitability, and the
prospects for future progress in searching for signs of life in the Universe as the outcome
of the NExSS workshop held in Nov 29-Dec 2, 2016, New Orleans, LA.
This is one of five "Life Beyond the Solar System" white papers submitted by
NExSS. The other papers are: (1) Exoplanet Astrophysical Properties as Context for
Habitability; (2) Technology Development Required for Future Progress; (3) Remotely
Detectable Biosignatures; (4) Observation and Modeling of Exoplanet Environments.
1. Areas of significant scientific or technological progress since publication of the
NASA Astrobiology Strategy 2015
From the perspective of ESW, major developments since AS15 are the following:
A. Exoplanet Observations
1. Discovery and characterization of superflares on K-M dwarfs, their frequency and
2. Observational search for CMEs from active stars has recently started [7,8].
3. Detection and characterization of exospheres in hot Jupiters and constraints on star-
relations to spot sizes, rotation and effective temperatures [4-6].
planet interaction (X-ray and Extreme UV (XUV) driven evaporation) models [9].
4. Characterization of XUV fluxes from K-M dwarfs using combined HST, EUVE,
Chandra and XMM-Newton data [10,11].
5. Reconstruction of Zeeman Doppler Imaging in a number of G-M dwarfs as a
prerequisite to constrain space weather models [12,13].
6. Detection of radio emission from substellar objects, extending down to a mass of
12.7 +/- 1 MJup, confirming magnetic field strengths >3000 G for the latter [14].
7. Development of the capability to conduct near-continuous simultaneous monitoring
of 1000s of nearby systems for radio emission (stellar CMEs, planetary auroral
emissions) and optical emission (stellar flares).
B. Modeling of Stellar and Planetary Environments
1. 3D magnetohydrodynamic (MHD) multi-fluid models of stellar winds and CMEs
have recently been constructed using advanced data-driven MHD tools validated and
calibrated for solar wind models [2,15-19]. These simulations suggest that fast, dense
winds and powerful CMEs disturb exoplanetary magnetospheres, generate ionospheric
currents, and introduce a number of effects including electron precipitation and Joule
heating. These effects need to be characterized to build a comprehensive picture of
their impacts on atmospheric erosion, particularly for HZ planets orbiting M
dwarfs which will be the first targets to characterize Earth-like exoplanets.
2. 1D multi-fluid coupled hydrodynamic and kinetic models of XUV driven ion
escape from exospheres of Earth-like exoplanets suggest that large XUV fluxes from
active planet hosting M dwarfs stars may contribute to atmospheric erosion on geological
timescales thus making exoplanets within their HZs uninhabitable [2,3]. Determining the
timescales over which these stars are active and the extent of atmospheric erosion is vital
for understanding exoplanet characterization and target selection with JWST.
3. 1D photo-collisional models enhanced with neutral chemistry were recently
applied to model the prebiotic chemistry driven by precipitation of energetic protons due
to SEPs from the young Sun and active stars [2,16,20].
C. Technology
1. Development of direct imaging techniques in the mid-infrared (IR) bands with Exo
type stellar
(extended Fourier-Kelvin
Life Beacon Space Telescope, ELBST
interferometers (FKSI) mid-IR space interferometers).
In the upcoming decade the exoplanet and astrobiology communities need to
prepare and develop future mission concepts for space interferometry missions to directly
image exoplanets in the near- and mid-IR around nearby solar type stars. The IR spectral
region (3-28 microns) is well known for its richness of molecular features from bands of
molecules such as carbon dioxide, water vapor, nitrous oxide, methane, hydroxyl and
nitric oxide. Considerable technology development for mid-IR nulling interferometers
began with the Keck Interferometer Nuller (KIN), and recently the LBTI that have
provided the most sensitive observations to date of the luminosity function of warm
debris disks in the HZs of nearby solar type stars. Testbeds for space interferometers
(TPF-I/Darwin/FKSI) have also been developed in the US and Europe.
2. OST development.
The Origins Space Telescope (OST) is one of four mission concepts currently
being studied by NASA in preparations for the Astrophysics 2020 Decadal Survey. It
features a large (6.5 - 9 meter), cold (4 K), mid-to-far-IR telescope that will be orders of
magnitude more powerful than existing facilities. OST will address this key science
question by characterizing the atmospheres of Earth-size planets transiting in the HZs of
mid-to-late M dwarf stars. OST will expand on the legacy of exoplanet science by
obtaining high-precision transmission and emission (dayside and phase-resolved) spectra
from 5 - 25 microns at a resolution R = 100 – 300. Achieving the necessary precision
with this proven technique requires the design of a purpose-built instrument. Continued
development of detector technology in the mid-IR is a fundamental step for the detection
of biosignatures in exoplanetary atmospheres.
2. Important scientific or technological topics omitted from the NASA Astrobiology
Strategy 2015 and which have seen advancement since publication of the strategy
Following the progress in our understanding space weather impacts on the Earth and
Mars due to recent missions (GRACE, CHAMP, MAVEN), the exoplanetary community
[22] has initiated development of new approaches omitted from the NAS15 to
characterize the impacts of ESW on close-in exoplanets around M dwarfs, including
Proxima-b and TRAPPIST-1 [2,3,15,21,22].
3. Key research goals in the search for signs of life in the next 20 years
A. Planet Hosting Stars:
1. ESW models for K-M dwarfs require the following observational inputs: i. Far UV,
ii. Physical parameters of stellar
Near-UV, XUV and radio emission fluxes;
chromospheres and coronae; iii. Surface magnetic field distribution (magnetograms).
2. Observed magnetic structures including spots and their association with flares.
3. Refine characterization of stellar ages based on a set of observables including Li,
rotation, CaII H&K, patterns of magnetic activity. Thus, dedicated observations of flares
on K-M stars at different phases of evolution are required along with flare frequency.
B. Star-Planet Interactions:
1. Develop coupled MHD, hydrodynamic and kinetic models that describe the
coupling of energy flows of planet-hosting stars, and their dissipation in magnetosphere-
mesosphere exoplanetary environments. This requires a well-coordinated and funded
interdisciplinary effort from HAPE community.
2. Derive thresholds on parameters of space weather from stars to make a planet
habitable (atmospheric neutral and ion escape rates).
3. Characterize chemistry changes due to: FUV, XUV, stellar winds, & particles.
4. Search for radio and optical stellar CME signatures by performing extended long-
term observations at lower frequencies (< 10 MHz) with space or lunar radio missions.
5. Search for planetary outflows in spectral lines of H (hot Jupiters) and nitrogen and
metals (terrestrial planets) driven by powerful stellar flares from active K-M dwarfs.
6. Explore when M dwarf habitable cases actually shift beyond the ice line due to
severe ESW, when combined with ameliorating internal heating, including radiogenic
sources as well as tidal heating within compact multi-body TRAPPIST-1 analog systems.
C. Exoplanet Environments:
1. Explore how ionosphere-thermosphere systems respond to extreme space weather.
2. Search for N2 through mid-IR transmission and direct imaging observations, as
necessary to determine how common N2 is within exoplanetary atmospheres.
3. Detect the chemistry of young terrestrial-type exoplanets "pregnant" with life:
signatures of prebiotic chemistry.
4. Detect signatures of hydrogen-rich (primary atmospheres) of terrestrial-type
exoplanets around very young planet hosting stars.
5. Understand exoplanet magnetic dynamos, mantle activity, and the interplay
between volcanic/tectonic activity and the generation of Earth-like magnetic fields.
6. Explore the role exomoons play in maintaining exoplanetary magnetic dynamos?
(e.g., tidal enhancement of convection vs. the possible tidal melting of inner cores.)
4. Key technological challenges in astrobiology as they pertain to the search for life
in extrasolar planetary systems
A. Direct Imaging
1. The Large Binocular Telescope Interferometer (LBTI) Hunt for Observable
Signatures of Terrestrial Planets (HOSTS) study has recently set new limits for exozodi
detection for solar-type stars [23]. These results demonstrate the power of LBTI for
vetting potential targets for future direct imaging missions such as LUVOIR or HabEx,
and the importance of completing and enlarging the study in the next few years.
2. Direct imaging techniques with FKSI-type ELBST.
Ground-based prototypes demonstrating relevant technologies and obtaining
important science were the Keck Interferometer Nuller, and the LBTI HOSTS project
[23]. Development of mission concepts and technologies were curtailed due to budget
issues in the last decade. However, recent studies of star-planet interactions, including the
interaction of coronal mass ejections with the atmospheres have shown that the
atmosphere of the Earth (viewed as an, NO, and other molecules [20, 24]. Exoplanetary
upper atmosphers respond strongly in the mid-IR and cools through mid-IR lines of NO
and CO2 and open a new potential of mid-IR spectroscopy of exoplanet atmospheres, not
only with OST, but also with future ground-based and space- or moon-based nulling
interferometers [25].
5. Key scientific questions in astrobiology as they pertain to the search for life in
extrasolar planetary systems
1. How can we detect spectral signatures of prebiotically important molecules
highlighting fundamental prerequisites of life including nitric oxide and nitrous oxide?
2. What chemistry of the most abundant and biologically important molecules that
participate in pathways producing complex sugars, amino acids, and nucleobases can be
learned from the biochemistry community studying origin of life on Earth?
3. How can astrophysics inform laboratory experiments in understanding which pathways
efficiently produce biologically important molecules?
4. What steps are needed to build a unified network of theorists, observers, and laboratory
scientists to explore the most efficient, laboratory validated, and calibrated methodologies
to characterize the biologically important molecules with the strongest spectral signatures
(high signal-to-noise, low spectral resolution) of life?
5. Can vibrant/detectable biospheres exist shielded from space weather in oceans below
ice shells, beyond the classical HZ (including icy moons and nomad/rogue worlds)?
6. Scientific advances that can be addressed by U.S. and international space
missions and relevant ground-based activities in operation or in development
1. TESS will greatly expand the population of known potentially habitable exoplanets,
some of which may be selected for characterization by JWST transit transmissions
spectra to look for signs of potential biosignature gases.
2. JWST will provide mid-IR transit and eclipse spectra of exoplanets around nearby
stars, particularly M and K stars with exoplanets discovered by TESS, allowing
characterization of their atmospheres.
3. ELTs and other ground-based platforms will greatly expand the list of rocky planets
orbiting ultracool stars and characterize the atmospheres of some of them.
4. WFIRST will demonstrate the coronagraph technology for a future direct imaging
mission that would study Earth-like planets, if total mission cost can be limited.
7. How to expand partnerships (interagency, international and public/private) in
furthering the study of life's origin, evolution, distribution, and future in the
Universe
1. NExSS's interdisciplinary community has an opportunity to formulate well-
defined complex questions that can be addressed using a systems approach. To enhance
the efficiency of this approach in searching for signs of life, we must also incorporate
Origins of Life/Biology methodologies into these studies.
2. The International Space Science Institute (ISSI) is an efficient model of scientific
collaboration in diverse fields of space science focusing on one fundamental challenge
[26]. International science conferences are another important avenue to highlight
challenges in searching for signs of life. We find that having only invited talks that set the
stage for breakout discussions has been a novel approach to foster collaboration. From
this perspective, the NExSS sponsored ESW workshop was a useful tool to connect and
unify an emerging community that brings diverse ideas and methodologies to the table.
3. The key element of collaborative efforts should be the inclusion and coordination
of international mission observations, theory, and laboratory experiments to explore
laboratory validated, and calibrated methodologies to find the strongest signs of life.
4. International structures should explore observational methodologies through their
national agencies with participation of public-private partnerships, such as the
Breakthrough Initiative. This foundation plans to develop a low-cost mission to help
search for life on Enceladus and its partnership with NASA can accelerate the project.
References:
1. Schrijver, C. J.; Kauristie, K. A., Alan D., Denardini, C. M. and 22 coauthors (2015) Advances in
LPI Cont. No. 1989.
25. Defrére, D., Hinz, P. M., Mennesson, B. et al. (2016), ApJ, 824, 66.
26. International Space Science Institute (ISSI) http://www.issibern.ch
Space Res., 55, 2745 (2015)
2. Airapetian, V. S., Glocer, A., Khazanov, G. V., Loyd, R. O. P., France, K., Sojka, J., Danchi, W.,
Liemohn, M. W. (2017) Astrophys. J., 836L, 3A.
3. Garcia-Sage, K., Glocer, A., Drake, J. J., Gronoff, G., Cohen, O. (2017) ApJ Let, 844, L13
4. Maehara, H. et al. (2012) Nature, 485, 478.
5. Maehara, H., Notsu, Y., Notsu, S. and 5 o-authors (2017) PASP, 69, 41.
6. Davenport, J. R. A., Kipping, D. M., Sasselov, D., Matthews, J. M., Cameron, C. (2016) 821, L31.
7. Osten, R., Crosley, M. K. eprint arXiv:1711.05113.
8. Villadsen, J., Hallinan, G., Bourke, S. (2016) Proc. of the IAU Symp., vol. 320, 191.
9. Lopez, E. D. (2017), MNRAS, 472, 245.
10. Loyd, R. O. P.; France, K., Youngblood, A. and 8 co-authors (2016) ApJ, 824, 102.
11. Youngblood, A., France, K., Parke L., R. O. and 18-co-authors (2017) ApJ, 843, 27.
12. Lüftinger, T., Vidotto, A. A.; Johnstone, C. P. (2015) ASSL, 411, ISBN 978-3-319-09748-0
13. Airapetian, V. S., Jin, M., Lüftinger and 3 co-authors (2018), submitted to Nature Astronomy.
14. Kao, M. , Hallinan, G., Pineda, J. S. and 3 co-authors (2017) AAS #229, id.408.06 .
15. Cohen, O. (2017) ApJ, 835, 220.
16. Airapetian, V. S., Glocer, A., Gronoff, G., Hébrard, E., Danchi, W. (2016) Nature Geoscience, 9, 452.
17. Vidotto, A. A., Bourrier, V. (2017) MNRAS, 470, 4026.
18. Lynch, B. J., Masson, S., Li, Y., DeVore, C. R. and 3 co-authors (2016) JGR 121, 10677.
19. Dong, C., Jin, M., Lingam, M.,Airapetian, V. S., Ma, Y, van der Holst, B. (2017), PNAS, doi: 10.1073/
20. Airapetian, V., Jackman, Mlynczak, M., Danchi, W., Hunt, L. (2017) Nature SREP. 7, article #14141.
21. Garraffo, C., Drake, J. J., Cohen, O., Alvarado-Gómez, J. D., Moschou, S. P. (2017) ApJ, 843, L33.
22. Airapetian, V. S. et al. (2018) Impact of ESW on Habitability, in preparation to Int. J. of Astrobiology.
23. Danchi, W., Bailey, V., Bryden, G. and 13 coauthors (2014) SPIE, vol. 9146, id. 914607.
24. Airapetian, V. S., Danchi, W. C., Chen, P. C., Rabin, D. M., Carpenter, K. G., Mlynczak, M. G. (2017)
pnas.170801011
|
1612.08380 | 1 | 1612 | 2016-12-26T13:36:43 | Short-Term Orbital Forcing: A Quasi-Review and a Reappraisal of Realistic Boundary Conditions for Climate Modeling | [
"astro-ph.EP"
] | The aim of this paper is to provide geoscientists with the most accurate set of the Earth's astro-climatic parameters and daily insolation quantities, able to describe the Short-Term Orbital Forcing (STOF) as represented by the ever-changing incoming solar radiation. We provide an updated review and a pragmatic tool/database using the latest astronomical models and orbital ephemeris, for the entire Holocene and 1 kyr into the future. Our results are compared with the most important database produced for studying long-term orbital forcing showing no systematic discrepancies over the full thirteen thousand years period studied.
Our detailed analysis of the periods present in STOF, as perturbed by Solar System bodies, yields a very rich dynamical modulation on annual-to-decadal timescales when compared to previous results.
In addition, we addressed, for the first time, the error committed considering daily insolation as a continuous function of orbital longitudes with respect to the nominal values, i.e., calculating the corresponding daily insolation with orbital longitudes tabulated {\it at noon}.
We found important relative differences up to $\pm$ 5\%, which correspond to errors of 2.5 W m$^{-2}$ in the daily mean insolation, for exactly the same calendar day and set of astro-climatic parameters. This previously unrecognized error could have a significant impact in both the initial and boundary conditions for any climate modeling experiment. | astro-ph.EP | astro-ph |
Short-Term Orbital Forcing: A Quasi-Review and a
Reappraisal of Realistic Boundary Conditions for
Climate Modeling
Rodolfo G. Cioncoa,∗, Willie W.-H Soonb
aComisi´on de Investigaciones Cient´ıficas de la Provincia de Buenos Aires – Universidad
Tecnol´ogica Nacional, Col´on 332, San Nicol´as (2900), Bs.As., Argentina
bHarvard-Smithsonian Center for Astrophysics, Cambridge, Massachusetts, 02138, USA
Abstract
The aim of this paper is to provide geoscientists with the most accurate set
of the Earth's astro-climatic parameters and daily insolation quantities, able
to describe the Short-Term Orbital Forcing (STOF) as represented by the
ever-changing incoming solar radiation. We provide an updated review and
a pragmatic tool/database using the latest astronomical models and orbital
ephemeris, for the entire Holocene and 1 kyr into the future. Our results
are compared with the most important database produced for studying long-
term orbital forcing showing no systematic discrepancies over the full thirteen
thousand years period studied. Our detailed analysis of the periods present
in STOF, as perturbed by Solar System bodies, yields a very rich dynami-
cal modulation on annual-to-decadal timescales when compared to previous
results. In addition, we addressed, for the first time, the error committed
considering daily insolation as a continuous function of orbital longitudes
∗Corresponding author
Email addresses: [email protected] (Rodolfo G. Cionco),
[email protected] (Willie W.-H Soon)
Preprint submitted to Earth-Science Reviews
June 27, 2018
with respect to the nominal values, i.e., calculating the corresponding daily
insolation with orbital longitudes tabulated at noon. We found important
relative differences up to ± 5%, which correspond to errors of 2.5 W m−2
in the daily mean insolation, for exactly the same calendar day and set of
astro-climatic parameters. This previously unrecognized error could have a
significant impact in both the initial and boundary conditions for any climate
modeling experiment.
Keywords: Astrometry and Geosciences, Orbital forcing, Solar irradiance,
Short-term perturbations
1. Boundary conditions for climate system: A brief historical re-
view of orbital modulations of incoming solar radiation
The historical development for a scientific understanding of weather and
climate change, particularly the astronomical theory of climate change is a
long one (Neumann, 1985). For example, the Roman writer on agriculture,
Lucius Columella (ca. 4 to 70 AD) began in his Book I De Re Rustica [On
Agriculture]:
"For I have found that many authorities now worthy of remembrance were
convinced that with the long wasting of the ages, weather and climate undergo
a change; and that among them the most learned professional astronomer,
Hipparchus, has put it on record that the time will come when the poles will
change position, a statement to which Saserna, [or Sasernas because J. Neu-
mann noted that this is most likely the writing from a father and son team]
no mean authority on husbandry, seems to have given credence. For in that
book on agriculture which he has left behind he concludes that the position
2
of the heavens had changed from this evidence: that regions which formerly,
because of the unremitting severity of winter, could not safeguard any shoot
of the vine or the olive planted in them, now that the earlier coldness has
abated and the weather is becoming more clement, produce olive harvests and
the vintages of Bacchus in the greatest abundance. But whether this theory
be true or false, we must leave it to the writings on astronomy."1
Fast-forward to modern scientific age, the well-known calls for "weather
forecasting as a problem in physics" and "climate as a problem of physics"
(c.f., Monin and Shishkov, 2000) as well as the recent quest for "a theory
of climate" (Essex, 2011) clearly exemplify and highlight the struggle of the
science of meteorology and climatology for the past century till today. We
may echo the most pristine quest set out by one M. Milankovi´c, among other
accomplishments also conferred as the father of climate modeling by Berger
(2012) upon a comprehensive review of the subject, that:
" . . . such a theory would enable us to go beyond the range of direct obser-
vations, not only in space, but also in time. . . It would allow reconstruction
of the Earth's climate, and also its predictions, as well as give us the first
reliable data about the climate conditions on other planets."2
Such facts and reality alone would demand an ever increasing scrutiny for
1The full text of this book is available at http://penelope.uchicago.edu/Thayer/E/Roman/Texts/
Columella/de Re Rustica/1*.html.
2From p. 6 of Petrovi´c (2009) where the first sentence of the quote has original words
taken directly from Milankovi´c's autobiography. We wish to note that the Wikipedia entry
for Milutin Milankovi´c https://es.wikipedia.org/wiki/Milutin Milankovi%C4%87
has incorrectly implied the quote to be from a 1913 paper entitled "Distribution of the
Sun radiation on the Earth's surface" (in Serbian).
3
a better precision and accuracy in defining and specifying what has been well
accepted to be the so-called external forcing boundary conditions for Earth's
complex and coupled climate system which involves not only the atmosphere,
land and ocean per se but must also necessarily includes a host of other fast,
slow and intermediate scales processes involving both the intrinsic magnetic
variations and modulation of the Sun's radiation outputs and the distinct
effects from gravitational interactions of all bodies in the Solar System (SS)
in affecting the geometrical variations of the Moon-Earth's orbital elements.
This basic requirement for a correct boundary condition for climate modeling
can be sharpened further if climate variations are considered as a characteris-
tic to be deduced from rather than artificially imposed on a realistic climate
model (e.g., Nicolis and Nicolis, 1995; Lions et al., 1997; Goody, 2007).
It is clear from a careful literature review (see detailed discussion be-
low for a brief review of those longer term orbital changes) that the limited
availability of STOF parameters has thus far prevented from a more direct
prescription of this particular boundary condition in any meteorological and
climatic simulations and studies of the past, present and future. To the best
of our knowledge, a rare exception is contained in the preliminary study by
Bertrand et al. (2002). A close inspection from the United Nations Fifth
Assessment Report (see Collins et al., 2013, p. 1051) tells us that STOF is
indirectly assumed to be unimportant and play no climatic role. The pri-
mary argument and assumption in neglecting changes in orbital forcing for
climatic changes over the last few thousand years, last century or even last
decade seemed to be from the claim that globally-averaged radiative forcing is
small or negligible. Other reasoning, involved in climate reanalysis projects
4
and products (see e.g., Simmons et al., 2014; Hersbach et al., 2015), seemed
to be that such an effect involved in the modulations of the seasonal forcing
and seasonality itself has been indirectly included in other boundary condi-
tions like through the prescription of sea surface temperatures or atmospheric
ozone concentration or through assimilation of other atmospheric metrics in
a 100-year atmospheric reanalysis simulation.
More than two decades ago, (Laskar et al., 1993, p. 525) echoed a re-
minder that:
"The mean annual insolation depends thus only on the eccentricity of
the Earth, and its variation over 1 Myr are . . . [very small, about 0.6 W m
−2 as shown in Figure 4 of Laskar et al. 1993] . . . In fact, this is not the
main paleoclimate quantity, and it was recognized by [Milankovi´c] . . . that the
summer insolation at high latitudes had a larger influence on the climate of
the past. If the insolation in summer is not high enough the ice does not melt,
and the ice caps can extend. This is why it is also important to compute the
daily insolation at a given point on the Earth."
We agree and only wish to call for a more direct accounting for STOF, as
a true boundary condition, in all climatic simulations in that the effects from
local and regional perspective are clearly not negligible nor unimportant (see
e.g., section 3.7 in Miller et al., 2014) in terms of seasonal dynamical evolution
of the coupled air-sea-land system. The importance of accurately specifying
the incoming solar radiation on local and regional geographic scales should
also be studied and assessed in contexts with the currently unresolved diffi-
culties in closing the budget of Earth's energy imbalance (Trenberth et al.,
2016) as well as in both interpreting and emulating local and regional surface
5
temperatures using the state-of-the-art general circulation models (Ji et al.,
2014; Lin and Huybers, 2016). Furthermore, the persistent modulation of
the orbital forcing, via the modulation of the seasonal irradiation ampli-
tudes, ranges and their geographical distributions, all across the globe may
yield a deeper insight to the basic problems of climate science on timescales
of decade to century.
We shall limit our exploration and calculation to the Holocene and 1 kyr
into the future, for which the longest available accurate ephemeris including
this period is the Jet Propulsion Laboratory's Development Ephemeris 431
(JPL's DE431) outputs (Folkner et al., 2014). Another simplification we
adopt for this paper is the postponement of the accounting of the intrinsic
variations of the Sun's irradiance outputs as caused by the variable nature of
solar magnetism on a host of timescales but a recent review on the choices of
Total Solar Irradiance (TSI) records to use for climatic study and modeling
has been reported in Soon et al. (2015).
We will not further speculate on the importance of getting the correct
boundary conditions for any atmospheric, meteorological and climatic sim-
ulations, but will limit ourselves to a brief note on some recent works as a
source of motivation. Cronin (2014) has recently highlighted the key role of
getting the solar zenith angle correctly represented in global climate models
where errors in assuming daytime-average zenith angle or spatially constant
insolation may lead to underestimating or overestimating of global energy
budget on the order of 10 W m−2 or more, respectively. Hogan and Hira-
hara (2016) documented the situations when the infrequent calls to radiation
module in a climate model calculation may lead to errors in stratospheric tem-
6
peratures of 3 to 5 K. He et al. (2015) spelled out the detailed dependence
of the amount of solar radiation on the length of day and latitude which in
turn for defining the role of Tibetan and Iranian Plateau in constraining the
northerly intrusion into the Indian monsoon area during boreal summers.The
multi-scale roles of solar radiation on the diurnal cycles and related coupled
ocean-atmosphere dynamics of the warm surface ocean including tropical In-
dian Ocean and the western equatorial Pacific have been examined by Shin-
oda (2005) and Li et al. (2013). Hudson et al. (2016) highlighted the key role
by solar irradiation impacts for the reflection heights in the ionosphere that
are changing by 15 km, 12 km and 1 km, respectively, by diurnal, seasonal
and 11-yr cycles of solar irradiation. Jajcay et al. (2016) shows and discusses
the evidence on how the small-amplitude 7-8 yr oscillations in the European
surface air temperature records can produce the so-called cross-scale modu-
lation in the temperature at a higher frequency range and regime, in their
case the annual cycles. It is far from clear that the multi-scale STOF reality
presented in our paper would have absolutely no role for this case study by
Jajcay et al. (2016). Finally, Joussaume and Braconnot (1997) and Chen
et al. (2011) have cautioned earlier on concerning the differences in calendar
assumptions (i.e., differences in the Sun's positions with respect to calen-
dar days) that can lead to a rather significant impact in climate modeling
through phase shifts on the prescribed solar irradiance values. This is why,
as emphasized in this work, an accurate and reliable way of tracking time in
the orbital calculation is important.
7
2. A quasi-review and overview of several important previous works
From a paleo-climatic perspective, the most widely applied studies re-
lated to changes in the incoming solar radiation (the so called insolation)
due to variations in both Earth's orbit and Earth's celestial pole, have been
considered in detail by, e.g., Berger (1978) and references therein, Berger and
Loutre (1991), Laskar et al. (2004, 2011a) for a very long time-span (i.e., ge-
ologic eras). Such long-term forcing, varying on tens to hundreds thousands
of years on incoming solar radiation, is often called the Milankovi´c orbital
forcing. These works are effort based on celestial mechanics publications of
Milankovi´c (1941), Sharaf and Boudnikova (1967), Bretagnon (1974), Berger
(1978), Laskar (1988), and Laskar et al. (2004) among others. Both Imbrie
(1982) and Berger (2012) proffered a comprehensive review on the history of
the orbital solutions which now have improved techniques and methodologies
starting with the classical orbital problem of the evolution of the gravita-
tional solar N -body system, which have their roots in the works of Laplace
and Lagrange (see e.g., Laskar, 2013).
These solutions (as it is usually referred to in astronomical literature for
the process in obtaining orbital elements and Earth's orientation parame-
ters from a semi-analytic model of the SS evolution) are intended for several
million years. Such calculations are based on the numerical integration of
the equations which describes the secular (i.e., long-term and long-period)
dynamics of the SS, where short-term periodicities, arising from the evolu-
tion of the orbital longitudes itself, were simply not taken into account in
order to reduce the complexity of the orbital integrations. Specially, the
works by Berger and Loutre (1991) cover the field for −5 Myr (i.e., before
8
present); Laskar et al. (2004) integrate a secular-realistic model of the SS,
given accurate solutions for −50 Myr and +20 Myr from present. Laskar
et al. (2011a) also using very accurate initial conditions and very short time
step (less than one year), have recently extended the previous solution to
−250 Myr, i.e., including all the Cenozoic era.3 These solutions coming from
works by Laskar and colleagues although permitting the insolation quantities
at, e.g., annual timescale, but all those impressive results did not account for
the STOF variations.
The key issue tackled in our present work is the problem of the short-
term variations of insolation quantities; i.e., the high-frequency orbital forc-
ing, and its variations at decadal and multidecadal scales. This problem was
first addressed by Borisenkov et al. (1983) and Borisenkov et al. (1985). They
determined short timescale variations between ∼ 2-20 yr which produce vari-
ations of the same order as the long-term orbital forcing in a few hundred
years. They found that the greatest short-term effect of orbital forcing occurs
in high latitudes at midsummer (July); with the lunar orbital retro-grading
period of 18.6 yr (i.e., the main nutation term) as the most evident driver
(see e.g., the recent evidence found for this modulation of sea level at the
Eastern North Sea and Central Baltic Sea regions by Hansen et al., 2015).
3In addition to what was already known about the chaotic evolution of orbital motions
of terrestrial planets in the SS with a characteristic Lyapunov time of 5 Myr (i.e., largely
limited by the lack of knowledge on the precise value of solar oblateness, J2), it is worth
noting that Laskar et al. (2011b) recently proposed that "it will never be possible" to track
the precise evolution of the Earth's eccentricity beyond 60 Myr owing to the strong chaos
caused by close encounters or even a chance collision between Ceres and Vesta.
9
This effect is found to be of the same order of amplitude as the intrinsic
irradiance variations due to solar magnetic cycle variations. Loutre et al.
(1992) performed the most complete analysis of short-term orbital forcing
to date. They analyzed the Earth's parameters involved in orbital forcing,
which are the so-called astro-climatic elements or parameters, for the last
6 kyr at an annual basis (i.e., getting one value for each calendar year).
Their astronomical solution is based on the French VSOP planetary theory
(Bretagnon, 1982). This semi-analytic theory, was developed in series up to
third order of the masses of all the planets considered plus an improvement
for the descriptions of the giant planets up to the sixth order. The lunar
perturbation in the Earth-Moon barycenter is also included; but the Earth
and the Moon are not considered as separated bodies. The basic integration
constants used were adjusted from former JPL DE200 ephemeris (Standish,
1982). Therefore, this solution allowed considerably shorter periodicities in
planetary perturbations than former astronomical solutions. Loutre et al.
(1992) confirmed the earlier results of Borisenkov and collaborators, showing
that the most important spectral peaks occur at decadal timescales (say, pe-
riods less than 40 yr), but also specified longer periodicities, some of them at
multicentennial to nearly millennial scales. They found that, for longer pe-
riodicities the precessional modulation produced the most important signals
for low to middle latitudes, with the obliquity signals stronger at solstices for
polar latitudes. Such an observation is clearly consistent with what is mostly
known for the Milankovi´c orbital timescales (see e.g., Imbrie, 1982).
It would be premature for us to spell out and prove the direct relevance of
all such STOF periodicities for any ranges of climatic measures and metrics
10
while the state of understanding and discussion in the literature remains
still largely unsettled. This is mainly because there has been little, with
the few exceptions reviewed here, study for STOF in terms of any actual
manifestations in the real external boundary conditions that can be subjected
to any systematic evaluations.
Although Milankovi´c forcing data are available (Berger and Loutre, 1991;
Laskar et al., 2004, 2011a), e.g., from the National Centers for Environmental
Information of NOAA/NCEI [formerly known as NOAA/NCDC]
(http://www.ncdc.noaa.gov/data-access/paleoclimatology-data/data
sets/climate-forcing),
detailed STOF data and the corresponding insolation quantities are still
largely missing from any scientific archives and literature. Secular (smooth)
values of insolation quantities can be obtained at centennial resolution us-
ing orbital solutions by Laskar and colleagues, from this web page (but, as
was mentioned, only strictly for Milankovi´c's scales of oscillations in tens of
millennia). To-date, there is no direct data made available publicly, e.g.,
from Loutre et al. (1992) solution, describing STOF and insolation quanti-
ties. Hence, we have decided to provide geoscientists with the most com-
plete and accurate descriptions of the STOF astro-climatic parameters using
DE431 ephemeris, the longest high precision ephemeris publicly available.
High precision ephemerides have the most accurate description of short-term
variations of Earth's orbit. Another important improvement (i.e., with re-
spect to Loutre et al. 1992) worth applying now is the use of the latest
precession-nutation formulas to get the astro-climatic parameters expressed
in a reference system of the date (i.e., with respect to the moving equinox).
11
Although a large effect may not be expected in our studied time-span of
the Holocene, it is important to use these latest formulations if an accu-
rate accounting of the dynamics of orbital forcing is desired. The classical
polynomial precession formulas (Lieske et al., 1977) are only valid for a few
centuries from J2000.0 (Vondr´ak et al., 2011) and non-rigid Earth nutation
series are now available (Mathews et al., 2002) for our inclusion.
From Milankovi´c works to present, the incoming solar radiation has been
reckoned in different ways (i.e., the so called insolation quantities:
instan-
taneous, daily, mean mid-month, monthly mean, seasonal, etc.). We think
that any arbitrary uses of insolation or derived quantities related to it, could
be confusing to expert modelers and non-expert readers alike. The basic
insolation definitions and calculations over arbitrary period of times are pe-
riodically reviewed and explained in terms of the most basic geometry or
refined mathematical tools (e.g., Berger, 1978; Laskar et al., 1993; Berger
et al., 2010). In this work we also feel that it is necessary to start review-
ing the basic definition of insolation. They are based on daily irradiation
over a whole Earth's variable rotation day (Berger, 1978; Berger et al., 2010;
Borisenkov et al., 1985; Loutre et al., 1992) which is the basis or foundation
of all longer time-span calculations.
Daily irradiation is the fundamental parameter of all insolation calcula-
tions and its formulation is derived considering the constancy; over a full day,
of astro-climatic elements and also the solar orbital longitude (therefore, the
solar declination; see Section 3), assuming that the error committed consider-
ing this constancy of orbital longitude is also negligible (Berger et al., 2010).
Nevertheless, for practical purposes, daily insolation formulas are usually ap-
12
plied, considering the metric as a continuous function of solar longitude; i.e.,
as a continuous function of time, setting a desired longitude value, which, in
fact, could be anything within a particular day. Such a procedure in adopt-
ing the daily insolation formulas as a continuous function of solar longitude,
is an habitual practice in paleoclimates studies, where insolation variations
at a particular orbital longitude (e.g., solstices, equinoxes, etc.) is followed
along the ages. The error with respect to the nominal (i.e., a constant solar
longitude) daily insolation formulas has not been addressed in the literature,
although one can assume a level of accuracy of 0.01 W m−2 in the mean
daily insolation, following the claim in Berger et al. (2010). In this work,
we have tabulated accurate daily insolation quantities at periods of one day
(one Julian day of 86400 s); so we can evaluate the difference or "error" in
evaluating the same insolation quantities, adopting both approaches, calcu-
lating daily insolation formulas as a continuous function of orbital longitude
and using tabulated values at noon of the corresponding day.
The main aim of this work is to describe and calculate the short-term
dynamics of the Earth's astro-climatic elements, with the highest precision
available and the basic daily insolation quantities suitable for climatic mod-
eling, for the whole Holocene up to 1 kyr in the future. We present our anal-
yses adopting the solar true orbital longitude coordinate in order to avoid
"calendar-problems" related to "fixed-day" calendar used in most paleocli-
mate studies (Joussaume and Braconnot, 1997; Chen et al., 2011; Steel, 2013).
Nevertheless, we also express our data in Julian days; which is a standard
time-reckoning in astronomy, this also provides a link to civil calendar of
the past and future. Although STOF variations produce departures of the
13
secular, long-term values less than ± 0.15 W m−2, we found that calculating
daily insolation quantities with tabulated values for a certain day or using
it as a continuous function of orbital longitude, can produces important dif-
ferences, up to ± 2.5 W m−2 (±5%) in the mean daily insolation, for the
same set of astro-climatic parameters. This holds even for the present day.
Again, along our theme of climatic variations should be a deduced character-
istics of the Earth's coupled non-linear dynamical system, we consider that
taking additional cares in accounting for this aspect of STOF boundary con-
ditions in any climate simulation or attribution studies are both necessary
and important.
3. Solar irradiance intercepted (or received) by the Earth and co-
ordinate systems for studying STOF variations
Let us define the instantaneous solar irradiance received at the top of the
atmosphere above a geographic location on Earth, at a certain moment of
a given date. The amount of solar energy flux, Qt, on a horizontal surface
per unit of time, depends on the received Total Solar Irradiance, TSI, (i.e.,
the total energy received from the Sun) which, as a flux, depends on the
inverse-squared-law to the source, for us the Sun-Earth distance r (reckoned
in astronomical units, au), and on the inclination of sun beams as follows:
(cid:18)1au
(cid:19)2
,
(1)
dQt
dt
= I = TSI(r) cos z; TSI(r) = TSI0
r
I is the instantaneous received solar irradiance, the intensity of the vertical
component of sunshine, or the incoming solar radiation (insolation); in MKS
system the insolation is reckoned in J/(s m−2) or W m−2. TSI0 is the "solar
14
constant", the solar irradiance received at a fiducial distance by a surface
directly oriented to the Sun. The angle subtended by the Sun from the
zenith, i.e., the zenithal distance, z, can be expressed as functions of the
latitude of the observer and of the declination of the Sun, δ, as shown in
Fig. 1. Here a local fixed geocentric equatorial system (X,Y, Z) is defined,
X directed towards the local meridian; −Z towards the hemispheric pole
(south, here); therefore the direction of the Sun, n(cid:12), and the direction of the
observer's zenith, nz, can be written as:
n(cid:12) = (cos δ cos H, cos δ sin H, sin δ); nz = (cos φ, 0, sin φ),
(2)
where H is the hour angle (reckoned from the local meridian, following the
sense of the Sun's apparent motion). To avoid confusion with sign conven-
tions, we assume that δ and φ are the absolute value of the celestial and
terrestrial angles to depict the problem and to consider in our formulation;
in subsequent applications we must put the corresponding signs (in fact in
Fig. 1, δ and φ need to be negative numbers to fulfill the right-hand rule of
the coordinate axes). Besides, we neglect the diurnal parallax effects, so the
geocentric angles are equal to topocentric ones (this assumption leads to a
correction at a maximum of 9 arcsec). It is thus straightforward that:
cos z = nz · n(cid:12) = sin φ sin δ + cos φ cos δ cos H,
(3)
then we can write the flux of the solar energy projected on a horizontal plane,
per units of time as:
dQt
dt
= I = TSI0
(cid:18)1au
(cid:19)2
r
(sin φ sin δ + cos φ cos δ cos H).
(4)
15
As a note, the "solar constant" for us is the index received at 1 au (a fix
distance to the Sun), e.g., 1364.5 ± 1.38 W m−2 (Mekaoui et al., 2010); but
an estimation based on the mean Sun-Earth distance: < r >= a(cid:112)(1 − e2),
where a is the Earth's orbit semi-major axis and e the Earth orbital eccen-
tricity; produces a different factor of (1 − e2)0.25 in Eq. (4). The importance
of getting the absolute value of TSI0 index correctly for climatology has been
recently discussed and explored in Soon (2014). For all the calculations in
this paper, we adopted TSI0 = 1366 W m−2 mainly to provide an apple-to-
apple comparison with the numerical results shown in Loutre et al. (1992)
but of course TSI0 is left as an adjustable free parameter in our computer
programs.
It is evident that for a fixed latitude, the insolation I depends strongly
on seasonal changes which are driven by annual solar declination variations,
with the Earth-Sun distance virtually constant at intra-annual timescale,
with respect to δ variations. This can be clearly seen by relating r and δ with
the Earth's astro-climatic parameters (Berger and Loutre, 1991; Loutre et al.,
1992); the above mentioned eccentricity, e; the longitude of the perihelion,
; and the obliquity of the ecliptic (Earth's obliquity), :
a(1 − e2)
1 + e cos(λ − )
r =
sin δ = sin(λ + 180◦) sin
;
(5)
where λ is the true longitude of the Earth on its orbit, a polar angle of
position of the Earth reckoned from the equinox of reference (γ0) which sets
the reference system on which these elements were originally depicted and
calculated (see Fig. 2). In what follows, we consider the Earth's semi-major
axis to be a non-constant parameter, because it has short-term variations
16
(but not secularly growing, at least, under the considered Holocene interval
of twelve to thirteen thousand years), hence this parameter will produce a
slight variation on the solar radiation received (of course, if we assume solar
constant evaluated at r = a; then a disappears in the formulation).
In paleoclimate studies, it is usual to set the Sun's longitude, which we
have already emphasized, and is defined as λ(cid:12) = λ + 180, and not the
Earth's longitude, because the apparent geocentric solar movement is de-
scribed. Hence, in order to identify the correct time of the year when sol-
stices and equinoxes occur, i.e., when the Sun passes through them in its
geocentric orbit, we have indicated in Fig. 2, the month of the year (for the
present calendar time) and the corresponding position of the Sun in its reflex
orbit. Here, the vernal equinox (origin of longitudes) corresponds to λ(cid:12) = 0
deg. (March), the June solstice is at λ(cid:12) = 90 deg. (i.e., the Earth is near its
aphelion), the equinox of September is at λ(cid:12) = 180 deg., and the solstice of
December λ(cid:12) = 270 deg. (i.e., when the Earth is near its perihelion).
At this point, it is important to remember that day by day, the origin of
orbital longitudes, the γ point, makes a complex movement involving com-
ponents of longer period (precession) and shorter period (nutation). Such
movements are produced by the combined planetary and luni-solar torques
which, in turn, affect the reckonings of longitude and also the Earth's obliq-
uity (see Fig. 2 and the next section); i.e., the Earth's pole positions on the
celestial sphere.
Eq. (5) clearly shows that the movement of the origin of longitudes has a
dramatic effect on the apparent positions of the Sun in the sky with respect
to the seasons. Because of this, we need to keep the synchronization of the
17
orbital longitudes with the seasons. Therefore, the origin of longitudes needs
to be measured from the "true –i.e, real and moving– equinox of the date" (γ);
hence if we call λt the true longitude and t the longitude of the perihelion
of the Earth measured from the true equinox of the date, we must replace
them in Eq.
(5); then the trigonometric part of both equations become
cos(λt − t), and, sin(λ(cid:12)t), with λ(cid:12)t as the true solar longitude of the date.
For the reference to calendar date (Gregorian, leap year mechanics), time is
kept following the precession, i.e., it is set, as far it is possible, to tropical
year, then the seasons are set with the months of the year at least for several
thousand years before and after present. Nevertheless, inconmensurabilities
between the involved periodicities and proper irregularities of the changing
Earth orbital elements, produces that fixed values of λt does not occurs
exactly at the same calendar date. In other words, the duration of the season
varies through time. In paleoclimate, a conventional calendar is used, it is the
present Gregorian calendar, with the present month versus Earth's longitude
relationship. Nevertheless, for the above-mentioned reasons (fixed values
of longitude do not occur exactly at the same date), important differences
in insolation estimates can occur when calendar date is used as temporal
reference, even with the conventional calendar (Joussaume and Braconnot,
1997; Steel, 2013, see below).
In what follows, we explain how to obtain high precision astro-climate
elements from DE431 for the studied time span. Because we used a standard
ephemeris, we obtain the basic data at a certain fix time step, based on a
fixed Earth's day of 86400 s, the Julian day. This is the standard way by
which the SS bodies are positioned at a certain moment. It is interesting to
18
note that this will also provide an exact link to the civil (Gregorian – Julian)
calendar.
4. Some astrometry
With the advent of fast computers, it is possible to resolve, by direct
numerical integration, the set of differential equations which describe very
accurate models of the planetary problem in the SS using high precision initial
conditions, for a very long time span. Moreover, the integrated orbits can be
fitted to observations which improves the final precision of the solution. In
this way, we can get for instance, the heliocentric position, r, and velocity,
v, of the Earth as a function of time; this process is referred to as the
ephemeris calculation. DE431 ephemeris models the Earth's center of mass
movement in the SS and some parameters of Earth's orientation in the inertial
space. The position and velocity of the Earth are obtained by adopting the
relativistic-barycentric reference system, related to the barycenter of the SS.
Later, for a specific moment, through an orbit calculation, we can replace
this set of six components of r and v with six "constants" of motion, the
Keplerian elements, which have the convenient property for describing the
orbital ellipses of the Earth around the Sun. Even more nicely, we can
describe the apparent geocentric orbit of the Sun around the Earth with such
a set of elements. Of course, through time, these "constants" vary because
there are more than two bodies (some of them significantly non-spherical
bodies) orbiting and interacting. Then, we need to follow through time the
Earth's orbital elements as well as the orientation of the Earth celestial pole
in space. These variations are the geometrical causes of insolation variations.
19
As mentioned, the longest high precision ephemeris publicly available is
the JPL's DE431 integration (Folkner et al., 2014). It is designed to cover
from 13,200 years before the reference epoch J2000.0 to 17,191 years after.
And DE431 is in good agreement with all the main ephemerides, i.e., the
French INPOP (Fienga et al., 2008) and the Russian EPM ephemeris (Pitjeva
and Pitjev, 2014). Unlike semi-analytical planetary theories such as VSOP
(Simon et al., 2013), numerically integrated ephemerides naturally include
all the terms of the disturbing functions; and integrates separately the Earth,
from the Earth-Moon barycenter so it provides the best description for short-
term variations on decade to century.
4.1. Issues of time keeping in brief
The time used in the dynamical equations of the ephemeris model is
reckoned in a relativistic (coordinate) timescale, expressed in TDB (time
dynamical barycentrical) units. This TDB time units evolves in a different
scale with respect to the timescale in which our civil time (the time expressed
in our watches) is based: the UTC (universal time coordinate). Our civil
time, can be easily related to true solar time, which describes the Sun's
position with respect to the observer's meridian at certain moment for a
date. But UTC is not a dynamical timescale (i.e., UTC is not a relativistic-
coordinate time). The corresponding coordinate time for an Earth observer
is TT (terrestrial time). Fortunately, the difference TDB−TT is less than 2
ms for several millennia (see e.g., Kaplan, 2006). So we can presumed TT
to be equivalent to TDB (and vice versa) as a temporal reference related
to orbital elements and all derived quantities from them. Therefore, we can
assume that one can accurately represent the Earth's orbital elements in
20
TT timescale. For example, a solar longitude recorded in DE431 at fiducial
J2000.0 epoch, corresponds to Julian day JD 2451545.0 (TT), or the calendar
date 1/1/2000, 12 h TT, which is equivalent to 1/1/2000 11:58:55.816 UTC;
i.e., at the approximate noon of an observer at Greenwich.
The relationship of UTC with TT has its own problems. UTC is based on
atomic international time, which has a uniform scale, but UTC is subjected
to leap second corrections. The corrections were done in order to keep UTC
nearest to the rotational time; which is in turn non-regular because of the
Earth's irregular rotation. The effect of the variability in Earth's rotation
has a clear documented evidence in the past and it has a fundamental role
on daily insolation determination, because the solar day depends on Earth's
rotation. Through this leap second correction, UTC is closely related to
the universal time (UT). This timescale has all the irregularities of Earth's
rotation. UT is studied and adjusted using ancient astronomical observations
and following Morrison and Stephenson (2004) for the interval 9999 BC-
700 BC, the difference TT−UT (which, for our purposes we can assume as
TDB−UT) follows a parabolic relationship such that for BC10000 = 1/1/-
9999 = JD−1931076.0 (TT), we have TT−UT = 5.17 h, which is often
referred to as the "clock error". This difference shows us that a certain
phenomenon (e.g., the Sun's position at noon) reckoned in TDB relativistic-
uniform timescale, has probably an uncertainty of that amount when we try
to refer as a geocentric phenomenon to be observed at that epoch. This also
implies that the present nominal 86400 s for the length of the day (LOD) is
no longer valid for ancient ages. This clock error is not usually mentioned nor
discussed in paleoclimatic studies, but it is important to take into account
21
for detailed modeling into the past. Therefore, the daily solar irradiance
that has been extensively used (Berger, 1988; Berger et al., 2010), based
on LOD of 86400 s, produces a very different estimation of the "real" daily
insolation received in a whole true solar day of a certain remote epoch in the
past (or future). Regarding civil calendars, it is important to remember that
calendar days refer to the Julian calendar dates before 4th October 1582.
Julian calendar has its own reckoning of the year duration; then, the JD for
some events in the past (as solstices or equinoxes) has an offset of several
calendar months with respect to the present calendar.
4.2. Obtaining Earth's orbital elements from DE431
The JPL website (ftp://ssd.jpl.nasa.gov/pub/eph/planets/Linux/de431)
provides a binary file of DE431 ephemeris, tailored for Lynux users, the
lnxm13000p1700.b file; the temporal limits of this file are 9/12/-13001 and
11/1/17000. As far as we know there is no simple program (e.g., FOR-
TRAN code) to read and directly obtain r and v from it. But JPL provides
a Fortran code for testing another version of ephemeris in binary blocks,
THESTHEP.f routine. We have modified this routine to get Earth's r and
v with full precision for an arbitrary time step. Then, we performed an
orbital calculation with a time step at every one Julian day. Thus, we are
getting astro-climatic elements at each day which is a huge improvement over
other estimates which considered only annual-resolution settings (e.g., Loutre
et al., 1992). In this way we obtain a full set of Earth's orbital elements and
positional angles: the semi-major axis, eccentricity, inclination, longitude of
the perihelion, longitude of the ascending node, mean anomaly, true longi-
tude and the true anomaly all with respect to the reference epoch J2000.0, in
22
TDB (cid:39) TT timescale. Therefore, for present times or instances, assuming a
constant Earth rotation speed (86400 s), an observer at Greenwich meridian
has such a precise set of Earth's elements and solar longitudes, at noon, to
calculate insolation quantities.
In what follows, and mainly to represent quantities as function of the
time, when we make mention of "time in years", for a specific Julian day
number J, we refer to Julian epoch:
t = 2000.0 +
J − 2451545.0
365.25
,
(6)
which enable us to reckon the variable time, t, as a real number (i.e., in years
and the fraction of a year).
The orbital calculation only determines and λ measured from the fixed
fiducial (γ0) J2000.0 mean equinox and ecliptic. However, in order to obtain
these quantities but measured with respect to the true equinox of the corre-
sponding date, we must deal with the long-term and short-term motions of
both the equator and the ecliptic.
5. The motion of basic planes
Luni-solar and planetary torques modify the Earth's celestial pole and
the pole of the ecliptic, which are described by the combined actions of the
longer term (i.e., longer than 100 centuries) precessional as well as the shorter
nutational changes. Mean precessional changes describe the evolution of the
mean equator and ecliptic at a certain date. In our STOF calculation, we
want to take into account the shortest variations of these planes; i.e., we
need to describe both the true ecliptic and equator of the date. The eclip-
23
tic is defined as the mean plane of the Earth-Moon barycenter orbit around
the SS barycenter (i.e., perpendicular to the Earth-Moon orbital angular
momentum vector with respect to the SS barycenter). Both, the ecliptic
and equatorial planes, are subjected to luni-solar and planetary perturba-
tions (Vondr´ak et al., 2011; Capitaine and Soffel, 2015). Fig. 3 shows the
precessional elements with respect to the J2000.0 reference epoch. The accu-
mulated precession (subscript A) of the equator with respect to the reference
epoch is described by ψA and ωA (in longitude and obliquity, respectively);
the corresponding elements of the ecliptic precession are ΠA and πA. The
mean obliquity of the date (i.e., only affected by precession) is A. Then, the
general precession in longitude accumulated from J2000.0 is:
pA = ΛA − ΠA,
(7)
an arc of great circle describing the absolute motion of the mean γ point along
the ecliptic of date; it combines the precession in longitude of equator and the
precession of ecliptic. To represent the long-term variations of these elements
we used the model provided in Vondr´ak et al. (2011), which is a significant
improvement with respect to the model described in Loutre et al. (1992) that
was in turn based on the older IAU standard of Lieske et al. (1977). Vondr´ak
et al. (2011)'s parametrization is based on numerical representations of mean
equator and ecliptic (e.g., Laskar et al., 1993), which are consistent with the
JPL's Development Ephemeris.
24
5.1. Combining secular and short-term variations for new astro-climatic el-
ements
The elements coming from DE431 define and orient the plane of the
Earth's orbit of the date with respect to the reference epoch (J2000.0). Recall
the fact that DE431 separates the Earth from the Moon, therefore we can de-
scribe the "true" Earth's orbit around the Sun. Then, for our astro-climatic
short-term solution, we need to combine the long-term/short-term variations
represented by precessional-nutational parameters and those specific Earth's
orbital orientation elements which contain both secular and short-term vari-
ations that in turn were derived from the DE431 ephemeris. This particular
geometry is described and summarized in Fig. 4. Although DE431 uses an
improved precession model for the orientation of Earth; which is a rigorous
numerical integration of the equations of motion of the celestial pole using
Kinoshita's solid Earth model (Kinoshita, 1977) for the speed of luni-solar
precession (Owen, 1990), which, however, is not the actual Vondr´ak et al.
(2011)'s model. We think that such a mismatch is not a big drawback because
the impact of Earth's true orientation on Earth's center of mass integration
should be small for the Holocene time-span we considered in this paper. We
contacted Dr William Folkner of NASA JPL and he strongly recommended
using Vondr´ak et al. (2011) precession formulas as the detailed model for
the long-term evolution of Earth's orientation to couple jointly to the DE431
ephemeris.
In Fig. 4, we replace the ecliptic of the day with the Earth's orbital plane,
defined through DE431, so we approximate:
25
πA = i; ΠA = Ω,
(8)
where i is the inclination of the Earth's orbit in the fiducial reference system;
Ω is the ascending node. In addition, we add the nutation in longitude and
obliquity to the equator movement, therefore:
ψt = ψA + ∆ψA; ωt = ωA + ∆ωA,
(9)
where ψA and ωa are obtained from Vondr´ak et al. (2011)'s theory. The
∆ψA and ∆ωA angles are the nutation in longitude and obliquity. The IAU
nutation model, based on the 1980 IAU theory, is now being replaced with
the new model of Mathews et al. (2002). Fortunately, its computational
implementation is provided in the Standards of Fundamental Astronomy
(IAU SOFA Board, 2016) through the SOFA software. We used IAU2000A
model, which is the most complete version that includes all planetary and
luni-solar effects on a non-rigid Earth with multiple couplings (see Mathews
et al., 2002).
At this juncture, it is important to note that the arc describing the moving
equinox (γ) over the orbital plane of the Earth is pt. We could call pt "the true
general precession of the date" because it is the general precession affected by
nutation and also by the approximate Earth's orbit as specified from DE431
ephemeris. Therefore, the arc N γ is now pt + Ω in this new description (see
Fig. 4). Hence, the changes in orbital longitude and in the longitude of the
perihelion, described in Section (3), should be written as:
t = + pt; λt = λ + pt,
(10)
26
and this is equivalent to a change in the origin of longitudes from γ0 to
γ in Fig. 2. Also the obliquity in Eq.
(5), must be replaced with the
"true obliquity of the date", t. This permits us to find a new astro-climatic
solution that we called the DE431-IAU solution.
6. New STOF astro-climatic elements: DE431-IAU solution
Now, we must find pt and t values as a function of time to obtain the
true longitudes and obliquity of the date. Consequently, we need to solve
the spherical triangle of Fig. 4, for which we follow a similar procedure as in
Loutre et al. (1992). Using cosine formula (by angle):
cos t = cos ωt cos i − sin ωt sin i cos(ψt + Ω)
(11)
and sine formula:
sin(pt + Ω) sin t = sin ωt sin(ψt + Ω)
(12)
and five elements formula (2 sides, 3 angles; see e.g., Duriez, 2002):
cos(pt + Ω) sin t = cos ωt sin i + sin ωt cos i cos(ψt + Ω)
(13)
we get:
tan(pt + Ω) =
sin ωt sin(ψt + Ω)
cos ωt sin i + sin ωt cos i cos(ψt + Ω)
.
(14)
We wish to note a typographical error in Loutre et al. (1992)'s Eq. (12):
the second member on the right hand side should be "+". By solving Eqs.
(11) and (14) we find a new set of astro-climatic parameters.
27
As a matter of consistency, we compare our results (pt and t); but re-
moving the effects of nutation correction, with the corresponding long-term
precessional elements of Vondr´ak et al. (2011), pA and A. The differences
are smaller than 30 arcsec in obliquity and smaller than 60 arcsec in lon-
gitude for the whole 13 kyr interval studied; which produce a very small
change in the solar radiation received, of ∼ 2 × 10−4 W m−2 at a maximum.
Such a difference is significantly smaller than the nominal error committed in
the theoretical calculation of daily irradiance (see next section). Therefore,
we argue that both DE431 elements and IAU related precession+nutation
models are sufficiently consistent and accurate for our purposes.
We show the full set of DE431-IAU astro-climatic elements (including
climatic precession, e sin ) over the full time interval in Figs. 5, 6, 7 and 8,
where one point for every 120-day has been plotted to avoid over-sampling
effects. We wish to note that these sampling effects can produce the appar-
ent spurious spikes in the representation of these elements, when you select
these data at a greater time step. Indeed, the smallest periodicities involved
in these orbital elements are of the order of a few months. The orbital peri-
ods of Mercury and Venus, which produce detectable spectral peaks in e.g.,
eccentricity, are about 0.24 yr and 0.6 yr, respectively (see also next section).
At first sight, and comparing with Loutre et al. (1992) solution, the DE431-
IAU solution seems to be "richer", with the evidence being most obvious
using the finest time resolution. The spectral analyses performed on them
confirm our claim.
28
6.1. Short-term periodicities detected
We have explored spectral analyses on the DE431-IAU solution by using
several methods; mainly, Lomb periodogram, maximum entropy and multi-
taper method (MTM), and have found a profuse amount of spectral lines
which basically include all the periodicities found in previous works. Short-
term periodicities in astro-climatic elements were reported by Borisenkov
et al. (1983) studying daily insolation variations at several latitudes; they
were: 2.7 yr, 4.0 yr, 5.9 yr, 11.9 and 18.6 yr. Loutre et al. (1992) determined
several short-term periodicities by analyzing first the variations of the astro-
climatic elements. For daily insolation at 65◦N, they found essentially the
same periodicities (showing that the most important spectral peaks occur
at decadal, say < 40 yr, time-scale), plus 8.1 yr, 15.7 yr and 29 yr periods;
arguing that these last were coming from perturbations by Venus, Mars and
Saturn. These are the main periodicities detected, but longer term periodici-
ties (e.g., ∼ 40, 60, 800, 900 yr) were also detected, but needing more detailed
scrutiny and careful examination. We agree with Loutre et al. (1992) that
MTM is a good choice for analyzing astro-climatic elements because this
method produces less noisy spectrum with an easy way to calculate confi-
dence levels; in this regard, we have used SSA-MTM toolkit from Ghil et al.
(2002). Our results show strong non-linear trends in the spectra (which is
clear from the element's graphics, Fig. 5-8), even while analyzing shorter
portions of the time series; therefore, our analysis include a prior detrending
of the time series. This is essential for a proper determination of the spectral
peaks corresponding to the longer periods.
In astro-climatic elements, these short periodicities are associated with
29
planetary mean motions or combinations of them because the Earth's dis-
turbing function depends on these orbital configurations (see Roy, 1978;
Murray and Dermott, 1999). These are generically called inequalities be-
cause the mean motion relationships p ni − q no, where p and q are integers
and ni, no are the mean motion of both planets, oscillates around a small
(say, few degrees per year) value, but are not exactly zero (the planets are
not in mean motion resonance; in fact, there are no exact mean motion res-
onances between planets in the SS). If the associated coefficient (Cj) in the
periodic term of the disturbing function is significant, then that term of the
perturbation is important. For example, if only two planets have been taken
into account, the periodic variations of Earth's orbital elements depend on
terms of the following form:(cid:88)
Cj
pni − q no
j
sin((p ni − q no) t + Bj)
(15)
where Bj is another coefficient depending on angular Keplerian elements.
Hence, if an inequality is small and the amplitude Cj is important, we are in
the presence of a significant perturbation, whose period is:
(cid:32)
(cid:33)
(16)
Per(p ni − q no) =
2π
ABS(p ni − q no)
= ABS
1
− q
Po
p
Pi
where Pi and Po are the orbital periods involved (note that p = q = 1 cor-
responds to the synodic periods of the pair of planets); ABS is the absolute
value function. The smaller the significant inequality, the longer the associ-
ated perturbation period. In the SS the most important interaction is the
Jupiter-Saturn (2:5) relationship, i.e., 2nj − 5ns (cid:39) −0.4 deg.
yr , which leads to a
30
Great Inequality of period ∼ 800-1100 yr (see e.g., Wilson, 1985); which was
detected by Loutre et al. (1992) in the evolution of eccentricity.
In order to identify the planetary origin of these short periodicities, we
need to evaluate the most important terms that produce detectable peaks in
the spectral analysis of the Earth's disturbing function, from Venus to, at
least, Saturn. The simplest and most complete way in achieving this is to
perform the high precision numerical simulations as provided by the MERCURY
software package for orbital dynamics (Chambers, 1999) and then evaluate if
any spectral peaks may be present (for a deep theoretical description and spe-
cific calculation of perturbations for a couple of planets the reader is directed
to Simon, 1987). First, we integrate the Earth's orbital motions adding only
one planet at a time from Mercury to Saturn, to identify the "pure" peri-
odicity of relevant perturbations; then we integrate the Earth with pairs of
"related" planets (Venus + Mars; Jupiter + Saturn; etc.) to see the modi-
fication to these "pure" periodicities, and then all the relevant bodies (from
Venus to Saturn) to see what, if any or all, of these short-term periodicities
"are surviving" despite the predatory effects from stronger combined pertur-
bations. By adopting this procedure we report in Fig. 9 that the prominent
peaks related to planet Venus are (in the order of importance, i.e., spectral
power):
3.98 yr = Per(2V-3E); 2.67 yr = Per(V-2E); 8.10 yr = Per(3V-5E); 7.84
yr = Per(5V-8E),
with V and E denote the mean motions of Venus and Earth. For the planet
31
Mars, the most conspicuous peak is:
15.76 yr = Per(E-2Ma),
where Ma denotes the mean motion of Mars. In Fig. 9, the minor peaks
labeled as P, P2 and P3, but with a confidence level less than 95%, come
from Mars and correspond to the following periods:
2.47 yr = Per(2E-3Ma); 2.9 yr = Per(3E-5Ma); 5.25 yr = Per( 3(E-2Ma)).
Peaks provided by Jupiter and Saturn mean motions (Per(J) = 11.86
yr; Per(S) = 29.4 yr) are also clearly visible. Of course the peaks around
6 yr and 14.7 yr are the second harmonics of J and S. The result of these
short-term simulations are in good agreement with main result of previous
works. It is important to note that the magnitude of the power spectrum
cannot be taken as a direct indication of the perturbation intensity. Only
an approximate estimation can be deduced from the raw data; the spec-
trum depends on detrending, sampling, etc., hence the relative intensity of
two peaks can largely vary for the same phenomena, if the input data are
processed straightforwardly.
Fig. 10 shows our MTM analysis (with a parabolic detrending of the
raw, original data shown in Fig. 5) of the 13 kyr evolution of the record of
eccentricity. All the peaks shown in Fig. 9 are detected in the detrended
time series. Particularly, some peaks, such as the P, P2 and P3 are rein-
forced; raising above the 99% confidence level (note that P2 is immersed
32
in the power of the signal at around 2.7 yr). Some higher harmonics such
as 4J, 3U (i.e., forth and third harmonics of Jupiter and Uranus mean mo-
tions, respectively); P4 (the sixth harmonics of 8E-15Ma, ∼ 41 yr); and P5
(fourth harmonics of the synodic period of Saturn and Neptune of ∼ 36 yr),
also appear. In particular, the perturbation with periodicities at around 3
yr strongly disturbs the (V-2E) periodicity and shifts this peak to 2.71 yr.
Longitude of the perihelion or climatic precession shows basically the same
periodicities (not shown). Notably, a spectral peak related to 4E-7Ma, not
previously shown, is strongly present at around 3.56 yr. The other observed
peaks are harmonics of the orbital frequencies present or non-linear combi-
nations of the physical parameters' frequencies intervening in the definition
of eccentricity.
Fig. 11 shows the MTM spectrum of obliquity from the detrended data
record. Although a cubic-parabolic detrending have been applied to the
raw series shown originally in Fig. 8, a secular trend in the spectral power
is still present. This spectrum shows some common periodicities with the
eccentricity and the precession of the perihelion signals. Periods directly
coming from nutation are clearly detected, especially the period of lunar
node (MΩ), and the inequality 2E-3Ma, which comes from the direct part of
planetary perturbation on Earth equatorial bulge (Souchay and Kinoshita,
1996, 1997; Souchay et al., 1999). Other direct planetary effects on nutation
come from 4E-7Ma, and 3E-5Ma.
In particular, a periodicity of 8.85 yr
is clearly visible that is related to the second harmonic of Saturn-Uranus
synodic frequency (period of about 17.9 yr), which produces considerable
torque in the SS (see Cionco and Abuin, 2016). Nevertheless, this 8.85 yr
33
periodicity can also be related to the small perturbation of the Moon on
the second order terrestrial potential J3 (rigid Earth), with the argument
−lM +F +MΩ; where lM is the mean anomaly of the Moon and F = LM−MΩ,
where LM is the mean orbital longitude of the Moon.
In summary, all of the above-mentioned spectral peaks are by far the
most important periodicities lesser than one century. Our solution DE431-
IAU is more complete in covering all the periodicities that were previously
known and shown; and some of these periods are expressed in daily insolation
quantities which in turn should be studied for any relations to weather and
climate as well as all the underlying physical processes.
7. Daily irradiance calculations
The daily irradiation or daily insolation, Q, for a whole rotational day
of duration τ and the mean daily irradiance, W , at a given latitude, can be
obtained by using Eq. (4), integrating over one full Earth's rotation, being
the integral that is only significant when the Sun is over the horizon.
In
addition if we assume elements others than H to be unchanging over a whole
day, the mean daily irradiance is:
W =
1
2π
(cid:90) π
(cid:18)1au
−π
I(φ, δ, r, H) dH
(cid:19)2 2
(cid:90) Hs
= TSI0
r
2π
0
(sin φ sin δ + cos φ cos δ cos H) dH, (17)
where Hs is the sunset hour angle. Finally:
(cid:18)1au
(cid:19)2
r
W =
TSI0
π
(Hs sin φ sin δ + cos φ cos δ sin Hs),
(18)
34
reckoned in the same units as TSI0 (W m−2). In addition, Eq. (4) permits
one to find the daily amount of solar irradiance between the time of sunrise
(ts) and sunset (tp) which is:
(cid:18)1au
(cid:19)2(cid:90) tp
r
ts
Q = TSI0
(sin φ sin δ + cos φ cos δ cos H) dt,
(19)
if we assume t to be the mean solar-time timescale (i.e., we drop the difference
with true solar time); using the rotational relationship with H (1 mean day
of τ seconds):
H =
(t − to),
2π
τ
(20)
then:
(cid:18)1au
(cid:19)2(cid:90) Hs
r
0
Q =
τ
π
TSI0
(sin φ sin δ + cos φ cos δ cos H) dH,
(21)
for one solar day of τ = 86400 s; and assuming the constancy of the Sun
declination, Q becomes:
(cid:18)1au
(cid:19)2
(Hs sin φ sin δ + cos φ cos δ sin Hs),
(22)
Q =
86400
π
TSI0
r
which is expressed in J m−2, if TSI0 is in W m−2. Eq. (3) permits one to
find the value of the H at sunrise-sunset. When the Sun is on the horizon
(i.e., Sun at 90 deg. of the zenith) the dot product is zero, hence:
cos(Hs) = − tan φ tan δ.
(23)
35
Eq. (23) acts as a discriminant for sunrise-sunset. If we call D = − tan φ tan δ,
then the Sun rises and sets at certain latitude when:
− 1 ≤ D ≤ 1.
(24)
If D = −1 for a particular observer; the Sun is at the limit of rising and
setting; i.e., when Hs = ±180 deg.; the Sun becomes circumpolar for that
observer: no sunrise-sunset for the time-span during D < −1. Of course, this
occurs when φ and δ have the same sign (the Sun at the same hemispheric
position). Then if:
D < −1,
(25)
the Sun does not set (polar day). When the Sun's declination changes sign,
the winter season arrives, and Hs decreases, and for certain latitude, it is
virtually zero; then D = 1. Therefore, the Sun does not rise (polar night), if:
(26)
and insolation quantities are null. If D < −1, then the integrand in Eq. (17)
is significant over 0 to 2π; hence, the mean irradiance for the circumpolar
D > 1,
Sun is:
Wc = TSI0
= TSI0
(cid:18)1au
(cid:18)1au
r
r
Then, Eq. (19) becomes:
(cid:90) 2π
0
(cid:19)2 1
(cid:19)2
2π
(sin φ sin δ + cos φ cos δ cos H) dH
sin φ sin δ.
(27)
36
(cid:18)1au
r
Qc = TSI0
= 86400 TSI0
(cid:19)2 86400
(cid:18)1au
(cid:19)2
2π
r
(cid:90) 2π
0
(sin φ sin δ + cos φ cos δ cos H) dH
sin φ sin δ.
(28)
Therefore, using Eq. (23) as discriminant, we can calculate by means of
Eqs. (18, 22, 27, 28), these daily solar quantities, knowing all of the other
corresponding solar-terrestrial parameters for a certain date. As a matter of
consistency, we see that the averaged quantities can be directly obtained in
dividing the daily values by 86400 (W = Q/86400).
In the derivation of Eqs. (18, 22, 27, 28) we adopted a day-based ap-
proach, i.e., taking the longitude data tabulated (at noon) for that date.
This particular approach is the conventional procedure commonly used in
obtaining daily solar quantities (see e.g., Berger, 1988; Laskar et al., 1993;
Berger et al., 2010). Nevertheless, as we have hinted at the beginning of
this article, it is an usual practise in paleoclimate studies, to calculate Q (or
W ) by fixing the Earth/Sun position at an arbitrary moment; i.e., fixing an
arbitrary value of solar longitude inside a particular day; this is equivalent
to taking Eqs. (18, 22, 27, 28) as a continuous function of longitudes. For
example, to follow the daily irradiation at March equinox for a mid-latitude
observer, the Eq. (18) is used by setting λ(cid:12) t = 0 as a temporal reference
in Eq. (5), to find δ. This approach is understandable because at ancient
time we do not know these parameters "exactly" at noon. Moreover, we can
assume that the change of solar longitude inside a day does not significantly
affect the estimate of daily insolation (Berger et al., 2010). Nevertheless, we
consider the clarification and recognition of this assumption to be important
37
for all users of available databases of solar irradiation (see next Sub-Section
77.1).
Fig. 12 shows the daily mean irradiation at equinoxes (λ(cid:12) t = 0, 180 deg.),
using our DE431-IAU solution, over the indicated calendar time, at 65◦N. The
solar constant used is 1366 W m−2. Fig. 13 shows similar quantities as in Fig.
12 but for solstices; i.e., λ(cid:12) t = 90, 270 deg. Both calculations were performed
taking Eq. (18) as a continuous function of solar longitude. As were pointed
out by Borisenkov et al. (1983) and Loutre et al. (1992), the secular long-term
modulation by precession is evident for equinoxes, even at high latitudes. For
solstices the dominant signal is provided by the modulation of the obliquity
variations. For daily insolation calculations, we prepare a FORTRAN code that
calculates Q and W (Eqs. 18, 22, 27, 28) along the time for a specific solar
longitude and latitude on Earth. The code INSOLA-Q-W.for permits the
calculation of Q or W for the exact mid-day at the tabulated longitude of
a certain Julian day, or to set Q or W as continuous functions of the solar
longitude, which is the common practice in paleoclimate studies. Fig. 12
and 13 were made with this last strategy, i.e., by setting λ(cid:12) t with the desired
value (0, 90, 180 or 270 deg.) using Eq. (18) and (22), but taking the DE431-
IAU elements for the specific tabulated JD within which the corresponding
value, in our cases of equinoxes or solstices, occurs. Hence, these values were
calculated for the specific astro-climatic parameters of the corresponding mid-
day, but did not used the tabulated mid-day longitude, as daily insolation
should be calculated in theory. The difference between the desired longitude
and the tabulated mid-day value can reach ∼ ± 0.7 deg.
Fig. 14 shows W at the "mid-month" of July; i.e., λ(cid:12) t = 120 deg. In
38
keeping with the usual practice in paleoclimate, we are continuing to adopt
the present calendar defining mid-month moments at an angular step of 30
deg., from March (0 deg.)
to February (330 deg.). Then the calculation
was also performed taking W as a continuous function of λ(cid:12) t. The period
corresponds to the same interval considered in Figure 12a of Loutre et al.
(1992), from 1450 AD to 1950 AD. In fact, this permits a direct comparison
of our result with Loutre et al. (1992); and this comparison yields a very good
agreement. We do not know the precise TSI0 value adopted by those authors
(but we presumed it was about 1366-1367 W m−2 during solar activity min-
ima as hinted on p. 191 of Loutre et al. 1992), but this level of agreement
is reached with 1366 W m−2. The obliquity signal is prominently present,
with the two main periodicities clearly visible: the ∼ 20 yr oscillation due to
lunar retro-gradating motion (18.6 yr) and the strong ∼ 2-3 yr period.
With regards to other periodicities, we performed an extensive spectral
analysis on daily insolation outputs (using detrended time series), especially
focused at the shortest periodicities on biennial to 40 yr timescales (the
main periods present in astro-climatic elements). For a comparison with
Loutre et al. (1992), we show the results for July mid-month at 65◦N (Fig.
15). Several periods visible in the eccentricity, longitude of the perihelion
and obliquity signals are present, but also prominent harmonics and non-
linear interactions of frequencies present in the parameters that define daily
insolation. The shortest periods oscillate at about 2.42 yr and 2.69 yr and
correspond to the perturbed periods by 2E-3Ma and V-2E, respectively. At
around 3 yr there are several harmonics and 3E-5Ma (2.9 yr), 4E-7Ma (3.56
yr), 4J; etc. Then 3.98 yr (2V-3E); 5.26 yr (3(E-2Ma)); 2J; 7.88 yr (5V-8E);
39
8.09 yr (3V-5E); Per(J); Per(2 MΩ), Per(2S), Per(MΩ) and Per(S) complete
the most important periods present. For lower latitudes (e.g., 0 deg.), the
spectrum is essentially the same; the only change is the attenuation of the
obliquity signal; as expected because precession signal is dominant at tropical
latitudes. At these lower latitudes, the 18.6 yr period and its harmonics
appears smaller than, for example, the Jupiter signal.
Although our derived periodicities are similar to Loutre et al. (1992), we
found a richer set of periodicities at around 2-3 yr, 5 yr, and 8 yr, coming
from a more complete description of planetary perturbations performed in
DE431 and also the more complete model of nutation. Also, by performing
a much more detailed detrending and sampling, we can also produce similar
results as those discussed in Loutre et al. (1992) for the longest periodicities
covering oscillations on timescales from multicenturies to a millennium.
In addition, we have created another code MONTH-GEO.for designing to
calculate the geographical distribution of mid-month insolation for arbitrary
years (between −10 kyr, 3 kyr), in different latitudinal bands. An output for
the year 1650, near the beginning of the Maunder Minimum, can be seen in
Table 1, for the indicated latitudes. This code calculates W , as continuous
functions of λ(cid:12) t, i.e., at the corresponding longitudes for each mid-month of
the conventional calendar. All our codes are available with the publication
of this paper.
7.1. The effect of the non-constant longitude within a day
As discussed in details in the previous section, Q and W formulas were
derived assuming that we have astro-climatic elements and true orbital lon-
gitudes at solar noon (H = 0) and these values do not change significantly
40
over a complete 2π variation of H. This approximation is supposed to lead
to a maximum theoretical error of 0.01 W m−2 in the determinations of W
(Berger et al., 2010).
In our STOF calculations, we know the value of the true solar longitude
at mid-day and the astro-climatic elements for that day. Therefore, we can
estimate the difference or "error" in the estimation of corresponding solar
quantity between the exact tabulated value at mid-day (Q0 and W0) and
the estimation considering Q or W as a continuous function of λ(cid:12) t: the
difference, Q−Q0, W −W0 and also the "relative error" ratio, (Q−Q0)/Q0 =
(W − W0)/W0. As far as we know, these differences have not been discussed
nor addressed in the literature.
To describe these differences as a function of time, we have shown in
Fig. 16 and 17 (for 65◦N) the relative error ratios evaluated at equinoxes
and solstices for all the 13 kyr studied period. At the solstices, when the
Sun is stationary, the relative errors are small. Although there are values
as large as 0.2%, they are all mostly less than 0.1%. For the equinoxes the
relative differences are larger, reaching ± 1.2%. Evidently, we are confirming
that the characteristic of this relative error is dependent on solar longitude.
To explore this, we calculate this error as a function of longitude for every
10-degree interval. The result is shown in Fig. 18. We observe very large
relative differences of up to ± 5% around December solstice.
The biggest relative differences between nominal Q or W values and the
corresponding values as continuous functions of time, occur at the moment
when the Sun is near the extreme declination at opposite hemispheres. This
means that these big relative errors occur in the seasons when the insolation
41
values are near minimum (of course, the same occurs for a southern observer);
then, there are not so important differences on the amount of modeled re-
ceived energy at these moments, as we can see in Fig. 19, where W values
(in continuous and tabulated modes) for λ(cid:12) t = 250 deg. (i.e., November-
December), are depicted for the last 1 kyr. Nevertheless, as we can see in Fig.
20, this error for the mean daily insolation, can increase up to ∼ 2.5 W m−2
in absolute values. These largest errors occur after the March equinox (at the
beginning of the boreal spring) and at the end of the boreal summer. The
behavior of these differences depends strictly on the trigonometric part of
Eq. (18), which is a monotonically increasing function of declination, which
varies approximately between 7×10−3 and 1 between December and June, re-
spectively, for a northern observer. The smallest errors are occurring around
solstice of December because of the very small differences resulting from the
subtraction between two smallest values of this function. Conversely, slightly
larger errors with respect to December solstice, occur near solstice of June,
when the expression for the trigonometric function in Eq. (18) has its largest
values.
The characteristics of this error is similar as a function time, over the full
13 kyr interval studied. Of course, this error could lead to large differences
for initial conditions in a climatic modeling code especially when the climatic
system is evolving nonlinearly while encountering the chaotic regime of the
climate dynamic phase space. Therefore, the difference in evaluating Q or
W as continuous functions of longitudes or tabulated values, provides a clue
of the "uncertainty" associated with the correct incoming solar radiation
when the daily insolation value is basically a guess in connection to climate
42
modeling.
7.2. Comparison with the state of the art works by J. Laskar and colleagues
The longest database available for Milankovit´c orbital forcing is the one
provided by Laskar et al. (2004), and its extension to −250 Myr in Laskar
et al. (2011a). While this solution in astro-climatic elements is given for
−250 Myr from present, inherent problems to its chaotic evolution make it
strictly valid, in paleoclimate studies, over about −50 Myr and +20 Myr from
the present, which is the time-span covered by former Laskar et al. (2004)
solution. This French solution (available at NOAA/NCDC which redirect to
L'Institut de M´ecanique C´eleste et de Calcul des ´Eph´em´erides -IMCCE-) is
consistent with Berger and Loutre (1991) for the last 5 Myr where latter was,
in turn, based on the original Laskar (1986, 1988) results. This is the reason
we have considered only the solution calculated by Laskar and colleagues
which we have labeled here as La2010 solution, as the authors named their
last solution in the paper of the year 2011 for a comparison to our DE341-
IAU solution. The comparison also allows indirect assessments of Berger and
Loutre solutions. The web interface at IMCEE, permits outputs from −100
Myr to +20 Myr from J2000.0, but at a minimum timestep of 100 yr; hence
this is the maximal resolution of this data that we took into account, which
is fine enough for our purposes.
We compare mean daily insolation at July mid-month (i.e., λ(cid:12) t = 120
deg.) for 65◦N, i.e., as a continuous function of longitude, using the INSOLA-Q-W.for
code. Fig. 21 shows this calculation for 12 kyr, from year −10000.0 to 2000.0.
Whereas our results describe the short-term variations on W , La2010 solu-
tion describes the mean variations (to make our results clearly visible at this
43
scale, we have decided do not graph La2010 solution from −12 kyr to −10
kyr, which describes the mean values of our DE431-IAU solution). A zoom
between −10 kyr and −9.6 kyr from J2000.0, shows the excellent agreement
of our result with La2010, and the effects of STOF on W . To show the differ-
ences between continuous and tabulated values in our DE431-IAU solution,
we performed the same calculation, using tabulated values at mid-day. Fig.
22 shows these differences in detail for the past 1 kyr. There are differences
up to ± 0.3% which represents a maximum error of about 1.2 W m−2 for
this particular July mid-month calculation. Hence, a user that calculated
the mean daily insolation using La2010 solution, can have an error up to 1.2
W m−2 considering solar position at the noon time of the corresponding day,
for July mid-month.
As a final commentary, Loutre et al. (1992) had shown that the level of
disagreement of their solution with the original solution from Berger (1978)
can be as large as 10 W m−2 for 10 kyr to 5 kyr before 1950 AD; while the
disagreement is more tolerable at level of less than 1 W m−2 for 5 kyr to
present. Hence, our work presents a better agreement with respect to the
state of the art solution by Laskar et al. (2011a). Our STOF solutions do
not show evidence for any long-term growth in comparison to the La2010
solution; La2010 describes virtually the mean value of our STOF solution.
The departure of our (continuously-varying-longitude) solution with respects
to secular variations are indeed small, less than about ± 0.15 W m−2 over the
full Holocene interval covered in our DE431-IAU solution. As an example, for
this July comparison, the long-term La2010 value corresponding to −1 kyr
from J2000.0 (obtained from web interface) is W = 430.162 W m−2, keeping
44
three decimal figures; our DE431-IAU solution (in continuous mode), for
mid-July, i.e., with λ(cid:12) t = 120 deg. (which correspond to t = 1000.577 yr
of our solution, with a true mid-day longitude of λ(cid:12) t = 119.673 deg.) give
us W = 430.238 W m−2, a difference of less than −0.018%. For −10 kyr
from J2000.0, La2010 gives W = 468.978 W m−2, whereas our solution gives
W = 468.968 W m−2 (at t = −8000.2382 yr), a difference of ∼ 0.002%.
This independent confirmation suggests that our solution can be applied for
exploring any climate modeling experiments accounting fully for the STOF
aspects of the orbital forcings previously not available publicly.
8. Summary
We have presented a new set of calculations accurately accounting for
all short-term orbital modulations that can be considered an improved rep-
resentation of the boundary conditions relevant for any meteorological and
climatic studies over the Holocene interval and 1 kyr into the future. We have
reviewed the subject carefully and discussed all the steps and assumptions
involved in our calculations. Our new orbital solutions, DE341-IAU, offer
an internally self-consistent set of boundary conditions readily applicable for
climate model simulation, attribution and even assimilation studies. The
rich spatio-temporal dynamics of the persistent solar irradiation forcing of-
fers a realistic accounting of all STOF dynamics suitable for the evaluation of
all atmospheric, oceanic and coupled air-sea oscillations on timescales rang-
ing from diurnal, fortnightly, monthly, sub-annual, seasonal, annual, QBO,
ENSO, decadal to multidecadal timescales.
We offer, for the first time, a detailed determination of the error com-
45
mitted using daily insolation quantities as a continuously varying longitude
over the day (i.e., the most common practice in all paleoclimate modeling
study thus far) with respect to the nominal (based on solar position at noon)
values. We found errors up to 5% in daily insolation quantities, which cor-
respond to an absolute difference of 2.5 W m−2 in W calculations, which are
significantly larger than the theoretically expected error (i.e., 0.01 W m−2
reported in Berger et al., 2010) in the calculation of W . This error is of sim-
ilar order as the uncertainty detected by Joussaume and Braconnot (1997)
due to the "calendar effects" (i.e., 5-8 W m−2), which have been shown to be
considerably consequential on climate modeling (see also Chen et al., 2011).
However, when compared to the relatively well-documented calendar effect,
it is important to note that our continuous-versus-tabulated longitude error
is a different effect which can occur even in the same specific calendar day.
Therefore, climate modelers have to be aware of these differences and how
this particular error would spread and propagate along with time especially
when intra-annual insolation quantities are required. The sensitivity of initial
climate conditions to these differences should also be addressed.
Our next step will be to incorporate a range of estimates of intrinsic
changes in TSI for the past few thousand years as constrained both by solar
activity proxies from paleoclimatic evidence as well as observational study
of solar-type stars. The astro-climatic parameters database, the FORTRAN
codes and all the solar radiation values will be made available to any scien-
tists and climate modeling groups requesting for them. We will also make
effort to install our solar radiation forcing outputs on several popular data
archive centers including the one at NOAA/NCDC. Finally, after the com-
46
pletion of our calculation, we discover that another parallel and independent
effort on tackling STOF has been ongoing in Russia by Fedorov (2015) where
the author adopted the JPL DE 405/406 Planetary and Lunar Ephemerides
covering the time interval from 3000 BC to 2999 AD.
Acknowledgments
The authors acknowledge the support of the grant UTN-3577 "Efectos
Orbitales sobre la Irradiancia Solar Recibida Durante el Holoceno y los
Pr´oximos 3000 Anos" (2015-2016) of the Universidad Tecnol´ogica Nacional,
Argentina. The authors are indebted to William Folkner of NASA-JPL, for
an open and deep discussion of DE-JPL ephemerides models and details.
We also extend thanks to Jean-Louis Simon of Paris Observatory-IMCEE,
France, for sharing one of his papers and valuable commentaries. Special
thanks to John Chambers of Carnegie Institution of Washington, for sharing
his Julian Day routine that is especially useful for negative years. Finally, we
are grateful to Maria McEachern, the librarian at the John Wolbach library
of the Center for Astrophysics, for valuable help. W.S.'s work is indirectly
supported by SAO grant proposal ID 000000000003010-V101. W.S. would
also like to acknowledge the influence by Dr. Duncan Steel's independent
examination of the STOF question and to thank the late Bob Carter, Mi-
lan Dimitrijevic, Monika Jurkovics, Laszlo Kiss, Katalin Olah, and Dmitry
Sokoloff for helps and encouragements.
47
References
Berger, A., 1978. Long-term variations of daily insolation and Quaternary
climatic changes. Journal of the Atmospheric Sciences 35 (12), 2362–2367.
Berger, A., 1988. Milankovitch theory and climate. Reviews of Geophysics
26, 624–657.
Berger, A., 2012. A brief history of the astronomical theories of paleoclimates.
In: Climate Change. Springer, pp. 107–129.
Berger, A., Loutre, M. F., 1991. Insolation values for the climate of the last
10 million years. Quaternary Science Reviews 10, 297–317.
Berger, A., Loutre, M.-F., Yin, Q., 2010. Total irradiation during any time
interval of the year using elliptic integrals. Quaternary Science Reviews
29 (17), 1968–1982.
Bertrand, C., Loutre, M.-F., Berger, A., 2002. High frequency variations of
the Earth's orbital parameters and climate change. Geophysical Research
Letters 29 (18), 40–1–40–4.
Borisenkov, Y. P., Tsvetkov, A., Agaponov, S., 1983. On some characteristics
of insolation changes in the past and the future. Climatic Change 5 (3),
237–244.
Borisenkov, Y. P., Tsvetkov, A., Eddy, J. A., 1985. Combined effects of
earth orbit perturbations and solar activity on terrestrial insolation. Part I:
Sample days and annual mean values. Journal of the Atmospheric Sciences
42 (9), 933–940.
48
Bretagnon, P., Jan. 1974. Termes a longues p´eriodes dans le syst´eme solaire.
Astronomy & Astrophysics 30, 141–154.
Bretagnon, P., Oct. 1982. Theory for the motion of all the planets - The
VSOP82 solution. Astronomy & Astrophysics 114, 278–288.
Capitaine, N., Soffel, M., Aug. 2015. On the definition and use of the ecliptic
in modern astronomy. In: Malkin, Z., Capitaine, N. (Eds.), Journ´ees 2014
"Syst`emes de r´ef´erence spatio-temporels". pp. 61–64.
Chambers, J. E., 1999. A hybrid symplectic integrator that permits close
encounters between massive bodies. Monthly Notices of the Royal Astro-
nomical Society 304 (4), 793–799.
Chen, G.-S., Kutzbach, J., Gallimore, R., Liu, Z., 2011. Calendar effect
on phase study in paleoclimate transient simulation with orbital forcing.
Climate dynamics 37 (9-10), 1949–1960.
Cionco, R. G., Abuin, P., 2016. On planetary torque signals and sub-decadal
frequencies in the discharges of large rivers. Advances in Space Research
57 (6), 1411–1425.
Collins, M., Knutti, R., Arblaster, J., Dufresne, J.-L., Fichefet, T., Friedling-
stein, P., Gao, X., Gutowski, W., Johns, T., Krinner, G., Shongwe, M.,
Tebaldi, C., Weaver, A., Wehner, M., 2013. Long-term climate change:
Projections, commitments and irreversibility. In: Stocker, T., Qin, D.,
Plattner, G.-K., Tignor, M., Allen, S., Boschung, J., Nauels, A., Xia, Y.,
Bex, V., Midgley, P. (Eds.), Climate Change 2013: The Physical Science
Basis. Contribution of Working Group I to the Fifth Assessment Report
49
of the Intergovernmental Panel on Climate Change. Cambridge University
Press, Cambridge, United Kingdom and New York, NY, USA, p. 10291136.
URL www.climatechange2013.org
Cronin, T. W., 2014. On the choice of average solar zenith angle. Journal of
the Atmospheric Sciences 71 (8), 2994–3003.
Duriez, L., 2002. Cours de M´ecanique C´eleste classique. Laboratoire das-
tronomie de luniversit´e de Lilles 1.
Essex, C., 2011. Climate theory versus a theory for climate. International
Journal of Bifurcation and Chaos 21, 3477–3487.
Fedorov, V., 2015. Spatial and temporal variations in solar climate of the
Earth in the present epoch. Izvestiya, Atmospheric and Oceanic Physics
51 (8), 779–791.
Fienga, A., Manche, H., Laskar, J., Gastineau, M., Jan. 2008. INPOP06: a
new numerical planetary ephemeris. Astronomy & Astrophysics 477, 315–
327.
Folkner, W. M., Williams, J. G., Boggs, D. H., Park, R. S., Kuchynka,
P., Feb. 2014. The planetary and lunar ephemerides DE430 and DE431.
Interplanetary Network Progress Report 196, 1–81.
Ghil, M., Allen, M. R., Dettinger, M. D., Ide, K., Kondrashov, D., Mann,
M. E., Robertson, A. W., Saunders, A., Tian, Y., Varadi, F., Yiou, P.,
Feb. 2002. Advanced spectral methods for climatic time series. Reviews of
Geophysics 40, 3–1–3–41.
50
Goody, R., 2007. Maximum entropy production in climate theory. Journal of
the Atmospheric Sciences 64 (7), 2735–2739.
Hansen, J. M., Aagaard, T., Kuijpers, A., 2015. Sea-level forcing by syn-
chronization of 56-and 74-year oscillations with the Moon's nodal tide on
the northwest European shelf (eastern North Sea to central Baltic Sea).
Journal of Coastal Research 31 (5), 1041–1056.
He, B., Wu, G., Liu, Y., Bao, Q., 2015. Astronomical and hydrological per-
spective of mountain impacts on the Asian summer monsoon. Scientific
Reports 5.
Hersbach, H., Peubey, C., Simmons, A., Berrisford, P., Poli, P., Dee, D., 2015.
Era-20cm: a twentieth-century atmospheric model ensemble. Quarterly
Journal of the Royal Meteorological Society 141 (691), 2350–2375.
Hogan, R. J., Hirahara, S., 2016. Effect of solar zenith angle specification
in models on mean shortwave fluxes and stratospheric temperatures. Geo-
physical Research Letters 43 (1), 482–488.
Hudson, T., Horseman, A., Sugier, J., 2016. Diurnal, seasonal and 11–yr solar
cycle variation effects on the virtual ionosphere reflection height and im-
plications for the met office's lightning detection system, ATDnet. Journal
of Atmospheric and Oceanic Technology 33, 1429–1441.
IAU SOFA Board, 2016. IAU SOFA software collection.
URL http://www.iausofa.org
Imbrie, J., 1982. Astronomical theory of the Pleistocene ice ages: A brief
historical review. Icarus 50 (2), 408–422.
51
Jajcay, N., Hlinka, J., Kravtsov, S., Tsonis, A. A., Palus, M., 2016. Time
scales of the European surface air temperature variability: The role of the
7–8 year cycle. Geophysical Research Letters 43 (2), 902–909.
Ji, F., Wu, Z., Huang, J., Chassignet, E. P., 2014. Evolution of land surface
air temperature trend. Nature Climate Change 4, 462–466.
Joussaume, S., Braconnot, P., 1997. Sensitivity of paleoclimate simulation re-
sults to season definitions. Journal of Geophysical Research: Atmospheres
102 (D2), 1943–1956.
Kaplan, G. H., 2006. The IAU resolutions on astronomical reference systems,
time scales, and Earth rotation models. arXiv preprint astro-ph/0602086.
Kinoshita, H., 1977. Theory of the rotation of the rigid Earth. Celestial
Mechanics 15 (3), 277–326.
Laskar, J., Mar. 1986. Secular terms of classical planetary theories using the
results of general theory. Astronomy & Astrophysics 157, 59–70.
Laskar, J., Jun. 1988. Secular evolution of the solar system over 10 million
years. Astronomy & Astrophysics 198, 341–362.
Laskar, J., 2013. Is the solar system stable? In: Duplantier, B., Nonnen-
macher, S., Rivasseau, V. (Eds.), Chaos. Springer, pp. 239–270.
Laskar, J., Fienga, A., Gastineau, M., Manche, H., 2011a. La2010: a new
orbital solution for the long-term motion of the Earth. Astronomy & As-
trophysics 532, A89.
52
Laskar, J., Gastineau, M., Delisle, J.-B., Farres, A., Fienga, A., 2011b. Strong
chaos induced by close encounters with Ceres and Vesta. Astronomy &
Astrophysics 532, L4.
Laskar, J., Joutel, F., Boudin, F., 1993. Orbital, precessional, and insola-
tion quantities for the Earth from -20 Myr to +10 Myr. Astronomy &
Astrophysics 270, 522–533.
Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia, A., Levrard, B.,
2004. A long-term numerical solution for the insolation quantities of the
Earth. Astronomy & Astrophysics 428 (1), 261–285.
Li, Y., Han, W., Shinoda, T., Wang, C., Lien, R.-C., Moum, J. N., Wang,
J.-W., 2013. Effects of the diurnal cycle in solar radiation on the tropi-
cal indian ocean mixed layer variability during wintertime madden-julian
oscillations. Journal of Geophysical Research: Oceans 118 (10), 4945–4964.
Lieske, J., Lederle, T., Fricke, W., Morando, B., 1977. Expressions for the
precession quantities based upon the IAU/1976/system of astronomical
constants. Astronomy & Astrophysics 58, 1–16.
Lin, M., Huybers, P., 2016. Revisting whether recent surface temperature
trends agrew with the CMIP5 ensemble. Journal of Climate, in press.
Lions, J., Manley, O., Temam, R., Wang, S., 1997. Physical interpretation of
the attractor dimension for the primitive equations of atmospheric circu-
lation. Journal of the Atmospheric Sciences 54 (9), 1137–1143.
53
Loutre, M.-F., Berger, A., Bretagnon, P., Blanc, P.-L., 1992. Astronomi-
cal frequencies for climate research at the decadal to century time scale.
Climate Dynamics 7 (4), 181–194.
Mathews, P. M., Herring, T. A., Buffett, B. A., 2002. Modeling of nutation
and precession: new nutation series for nonrigid Earth and insights into the
Earth's interior. Journal of Geophysical Research: Solid Earth 107 (B4).
Mekaoui, S., Dewitte, S., Conscience, C., Chevalier, A., 2010. Total solar ir-
radiance absolute level from DIARAD/SOVIM on the International Space
Station. Advances in Space Research 45 (11), 1393–1406.
Milankovi´c, M., 1941. Kanon der erdbestrahlung und seine anwendung auf
das eiszeitenproblem (Canon of Insolation and the Ice-Age Problem). Acad.
Roy. Serbian, Editions spec (1969 translation by the Israel Program for
Scientific Translation; Edited by B. Benny and I. Meroz) 133, 1–484.
Miller, R. L., Schmidt, G. A., Nazarenko, L. S., co authors, Apr. 2014.
CMIP5 historical similations (1850-2012) with GISS ModelE2. Journal of
Advances in Modeling Earth Systems 6, 441–477.
Monin, A. S., Shishkov, Y. A., Apr. 2000. Reviews of topical problems: Cli-
mate as a problem of physics. Physics Uspekhi 43, 381–406.
Morrison, L. V., Stephenson, F. R., 2004. Historical values of the Earth's
clock error ∆t and the calculation of eclipses. Journal for the History of
Astronomy 35, 327–336.
Murray, C. D., Dermott, S. F., 1999. Solar system dynamics. Cambridge
Univ. Press.
54
Neumann, J., 1985. Climatic change as a topic in the classical greek and
roman literature. Climatic Change 7 (4), 441–454.
Nicolis, C., Nicolis, G., 1995. From short-scale atmospheric variability to
global climate dynamics: toward a systematic theory of averaging. Journal
of the Atmospheric Sciences 52 (11), 1903–1913.
Owen, W. M., 1990. A theory of the Earth's precession relative to the invari-
able plane of the Solar System. Ph.D. thesis, University of Florida.
Petrovi´c, A., 2009. Revolution and insolation: How Milutin Milankovi´c has
assembled the puzzle of the climate? Scientific Technical Review 59 (1),
3–10.
Pitjeva, E. V., Pitjev, N. P., Aug. 2014. Development of planetary
ephemerides EPM and their applications. Celestial Mechanics and Dy-
namical Astronomy 119, 237–256.
Roy, A. E., 1978. Orbital motion. IOP Publishing Ltd.
Sharaf, S. G., Boudnikova, N., 1967. Secular variations of elements of the
Earth's orbit which influences the climates of the geological past. Bull.
Inst. Teor. Astron. Akad. Nauk SSSR 11, 231–261.
Shinoda, T., 2005. Impact of diurnal cycle of solar radiation on intraseasonal
SST variability in the Western Equatorial Pacific. Journal of Climate 18,
2628–2636.
Simmons, A., Poli, P., Dee, D., Berrisford, P., Hersbach, H., Kobayashi, S.,
Peubey, C., 2014. Estimating low-frequency variability and trends in atmo-
55
spheric temperature using ERA-Interim. Quarterly Journal of the Royal
Meteorological Society 140 (679), 329–353.
Simon, J.-L., Mar. 1987. Computation of the first and second derivatives of
the Lagrange equations by harmonic analysis. Astronomy & Astrophysics
175, 303–308.
Simon, J.-L., Francou, G., Fienga, A., Manche, H., 2013. New analytical
planetary theories VSOP2013 and TOP2013. Astronomy & Astrophysics
557, A49.
Soon, W., 2014. Sun shunned. In: Moran, A. (Ed.), Climate Change: The
Facts 2014. Institute of Public Affairs, pp. 57–66.
Soon, W., Connolly, R., Connolly, M., 2015. Re-evaluating the role of so-
lar variability on northern hemisphere temperature trends since the 19th
century. Earth-Science Reviews 150, 409–452.
Souchay, J., Kinoshita, H., Aug. 1996. Corrections and new developments
in rigid Earth nutation theory. I. Lunisolar influence including indirect
planetary effects. Astronomy & Astrophysics 312, 1017–1030.
Souchay, J., Kinoshita, H., Feb. 1997. Corrections and new developments in
rigid-Earth nutation theory. II. Influence of second-order geopotential and
direct planetary effect. Astronomy & Astrophysics 318, 639–652.
Souchay, J., Loysel, B., Kinoshita, H., Folgueira, M., Feb. 1999. Corrections
and new developments in rigid Earth nutation theory. III. Final tables
"REN-2000" including crossed-nutation and spin-orbit coupling effects.
Astronomy & Astrophysics Supplement 135, 111–131.
56
Standish, Jr., E. M., Oct. 1982. Orientation of the JPL Ephemerides, DE
200/LE 200, to the dynamical equinox of J2000. Astronomy & Astrophysics
114, 297–302.
Steel, D., 2013. Perihelion precession, polar ice and global warming. Journal
of Cosmology 22, 10106–10129.
Trenberth, K. E., Fasullo, J. T., von Schuckmann, K., Cheng, L., 2016.
Insights into Earth's energy imbalance from multiple sources. Journal of
Climate, in press.
Vondr´ak, J., Capitaine, N., Wallace, P., 2011. New precession expressions,
valid for long time intervals. Astronomy & Astrophysics 534, A22.
Wilson, C., 1985. The great inequality of Jupiter and Saturn: from Kepler
to Laplace. Archive for History of Exact Sciences 33, 15–290.
57
Table LIST - CAPTIONS
Table 1: Mid-month W values at indicated latitudinal bands, for the
year 1650, at the beginning of the Maunder Minimum interval of reduced
sunspot activity; TSI0 = 1366.0 W m−2 (code MONTH − GEO.for). Each file
corresponds to one month, from March to February.
Table 1: Mid-month W values at indicated latitudinal bands, for the year 1650, at the
beginning of the Maunder Minimum interval of reduced sunspot activity; TSI0 = 1366.0
W m−2 (code MONTH − GEO.for). Each file corresponds to one month, from March to
February.
80◦S
65◦S
35◦S
0◦
35◦N
65◦N
80◦N
479.093
457.392
495.965
421.104
216.929
16.487
0.000
273.664
327.543
439.656
435.016
280.267
75.693
0.000
75.854
184.610
357.826
436.825
357.826
184.610
75.854
0.000
73.207
271.054
420.713
425.199
316.770
264.660
0.000
15.558
204.709
397.381
468.027
431.626
452.106
0.000
2.744
179.712
385.755
481.757
480.005
518.599
0.000
15.497
203.867
395.742
466.091
429.836
450.227
0.000
72.683
269.130
417.732
422.190
314.535
262.799
75.240
183.117
354.932
433.292
354.932
183.117
75.240
271.797
325.287
436.611
431.991
278.309
75.155
0.000
477.195
455.571
493.975
419.403
216.045
16.413
0.000
554.521
513.252
515.122
412.469
192.155
2.933
0.000
58
FIGURES LIST - CAPTIONS (ALL FIGURES IN BLACK-WHITE)
Fig. 1: A local fixed geocentric equatorial system (X, Y, Z) is defined;
X directed towards the local meridian; minus Z axis directed towards the
hemispheric pole (south, here). The minus Y axis points west. The positions
of the observer's zenith and the Sun are also indicated.
Fig. 2: Sketch of the Earth's orbit in the present time (reference epoch
J2000.0). The Sun (with the usual astronomical symbol (cid:12)) is at the fo-
cus. Nevertheless, to describe the apparent (also anti-clockwise), relative
movements of the Sun with respect to Earth, we have marked the Sun pro-
jections on the celestial sphere, specially for the corresponding solstices and
equinoxes, which permit a determination of λ(cid:12) to be used in the insolation
formulas. Also, the effect of the moving position of equinoxes (from γ0 of the
epoch to γ of the date) is marked with an arrow. The dotted segment q − A
is the apsidal line which determines the perihelia (q) and aphelia (A). The
direction to the ecliptic's pole is C. The Earth's equator is drawn over the
gray terrestrial sphere, and its continuation determines the Sun's declination
(δ) with the Sun-Earth radiovector (r).
In addition, the Earth terrestrial
pole direction is marked, which help defines the obliquity . The Earth's
true longitude and the longitude of the perihelion, both of the epoch of ref-
erence, are also indicated.
Fig.
3: Precessional elements measured with respect to the fiducial
J2000.0 epoch. The mean equator of reference (J2000.0) and the mean equa-
59
tor of the date, are indicated. The nodal point N is the intersection of the
two ecliptics: Eo of the reference epoch, and Et the ecliptic of the date. The
general precession in longitude accumulated from J2000.0 is pA (i.e., the seg-
ment from the projection of γ0 on Et, to γ point). See the main text for
descriptions of other angles.
Fig. 4: The spherical triangle used to find the "true general precession"
and obliquity of the date is shown. The Ω and i elements coming from DE431
define the plane of the Earth's orbit of the date with respect to the reference
epoch (J2000.0). See the text for more details.
Fig. 5: Short-term evolution of the Earth's eccentricity over 13 kyr (one
datum per 120 d). The last 500 yr, from 1950, is illustrated in the insert
with expanded details.
Fig. 6: Short-term evolution of the Earth's longitude of the perihelion
over 13 kyr (one datum per 120 d). The last 500 yr, from 1950, is illustrated
in the insert with expanded details.
Fig. 7: Short-term evolution of the Earth's climatic precession over 13
kyr (one datum per 120 d). The last 500 yr, from 1950, is illustrated in the
insert with expanded details.
Fig. 8: Short-term evolution of the Earth's obliquity over 13 kyr (one
datum per 120 d). The last 500 yr, from 1950, is illustrated in the insert
60
with expanded details.
Fig. 9: MTM power spectrum of the raw data from the simulated per-
turbations on Earth's eccentricity by Venus to Saturn over 1 kyr (from 1000
to 2000 AD). The labels mark the main recognized perturbations. P, P2 and
P3 refer to perturbations peaks with confidence level less than 95% which
are related to planet Mars. One un-labeled peak between P2 and 2V-3E
seems to be associated with the 4E-7Ma inequality. The other un-labeled
but significant peaks are harmonics of labeled perturbations.
Fig. 10: MTM power spectrum for the short-term eccentricity variations
based on the detrended data series after removing a parabolic long-term
trend. The basic periodicities identified in Fig. 9 can also be detected here.
Fig. 11: MTM power spectrum of the detrended obliquity time series.
The main periodicities are identified (see discussion in the main text). Other
periodicities are from harmonics and non-linear mixing of frequencies.
Fig. 12: Daily mean irradiation at 65◦N over the indicated calendar in-
terval for equinoxes; i.e., λ(cid:12) t = 0, 180 deg. The solar constant used was 1366
W m−2.
Fig. 13: The same as Fig. 12 but for solstices; i.e., λ(cid:12) t = 90, 270 deg.
The solar constant used was 1366 W m−2.
61
Fig. 14: W at July mid-month; i.e., λ(cid:12) t = 120 deg., for 65◦N. The time
interval corresponds closest to the one considered in Loutre et al. (1992)'s
Fig. 12a. This comparison shows a very good agreement with their results.
Solar constant used was 1366 W m−2.
Fig. 15: MTM spectrum of W (over detrended time series) at July mid-
month, 65◦N. The main periodicities are identified and are described in the
main text. At lower latitudes, the periodicities arising from the obliquity
signals are weakened.
Fig. 16: Relative difference in W calculation as a continuous λ(cid:12) t or using
tabulated values at mid-day, for solstices (λ(cid:12) t = 90, 270 deg.). At these mo-
ments, when the Sun is stationary, the errors are small, but they can reach
0.2%.
Fig. 17: The same as Fig. 16 but for equinoxes (λ(cid:12) t = 0, 180 deg.). The
error is larger on average by one order of magnitude (or about 1.2 %) when
compared to the errors during solstices.
Fig. 18: Relative difference in the W calculation as a function of λ(cid:12) t over
the 13 kyr interval we studied (for 65◦N). Around the solstice of December,
the relative error can reach 5% but drop to the level of less than 0.1% at 270
deg.
62
Fig. 19: W calculation as a continuous λ(cid:12) t (lines) or using tabulated
values at mid-day (dots), for λ(cid:12) t= 250 deg. (for 65◦N), from 1000 to 2000
AD. The absolute difference is less than 0.3 W m−2.
Fig. 20: Difference W −W0 in mean daily insolation calculation as a func-
tion of λ(cid:12) t for the 13 kyr interval we studied (for 65◦N). At boreal spring
and at the end of the boreal summer, the absolute difference can reach 2.5
W m−2 (see the text for more explanation).
Fig. 21: W at July mid-month (λ(cid:12) t = 120 deg.), for 65◦N: this work,
DE431-IAU (dashed lines); Laskar et al. (2011a) solution, La2010 (points).
The comparison is shown starting at −10 kyr with expanded details of the
intercomparison illustrated in the insert. The solar constant used was 1366.0
W m−2.
Fig. 22: W at July mid-month (λ(cid:12) t = 120 deg.), for 65 ◦N: DE431-
IAU solution, continuous values (dashed lines); La2010 data (points) and
DE431-IAU solution, with tabulated values (dots). Differences between the
continuous and tabulated values can reach ± 0.3%; i.e., approximately ± 1.2
W m−2.
63
Figure 1: A local fixed geocentric equatorial system (X, Y, Z) is defined; X directed to-
wards the local meridian; minus Z axis directed towards the hemispheric pole (south,
here). The minus Y axis points west. The positions of the observer's zenith and the Sun
are also indicated.
64
Figure 2: Sketch of the Earth's orbit in the present time (reference epoch J2000.0). The
Sun (with the usual astronomical symbol (cid:12)) is at the focus. Nevertheless, to describe the
apparent (also anti-clockwise), relative movements of the Sun with respect to Earth, we
have marked the Sun projections on the celestial sphere, specially for the corresponding
solstices and equinoxes, which permit a determination of λ(cid:12) to be used in the insolation
formulas. Also, the effect of the moving position of equinoxes (from γ0 of the epoch to
γ of the date) is marked with an arrow. The dotted segment q − A is the apsidal line
which determines the perihelia (q) and aphelia (A). The direction to the ecliptic's pole
is C. The Earth's equator is drawn over the gray terrestrial sphere, and its continuation
determines the Sun's declination (δ) with the Sun-Earth radiovector (r). In addition, the
Earth terrestrial pole direction is marked, which help defines the obliquity . The Earth's
65
true longitude and the longitude of the perihelion, both of the epoch of reference, are also
indicated.
Figure 3: Precessional elements measured with respect to the fiducial J2000.0 epoch. The
mean equator of reference (J2000.0) and the mean equator of the date, are indicated. The
nodal point N is the intersection of the two ecliptics: Eo of the reference epoch, and Et
the ecliptic of the date. The general precession in longitude accumulated from J2000.0 is
pA (i.e., the segment from the projection of γ0 on Et, to γ point). See the main text for
descriptions of other angles.
66
Figure 4: The spherical triangle used to find the "true general precession" and obliquity
of the date is shown. The Ω and i elements coming from DE431 define the plane of the
Earth's orbit of the date with respect to the reference epoch (J2000.0). See the text for
more details.
67
Figure 5: Short-term evolution of the Earth's eccentricity over 13 kyr (one datum per 120
d). The last 500 yr, from 1950, is illustrated in the insert with expanded details.
68
0.015 0.016 0.017 0.018 0.019 0.02 0.021−120−100−80−60−40−20 0EccentricityTime (Julian centuries from 2000.0)−4−3−2−1Figure 6: Short-term evolution of the Earth's longitude of the perihelion over 13 kyr (one
datum per 120 d). The last 500 yr, from 1950, is illustrated in the insert with expanded
details.
69
−100−50 0 50 100 150−120−100−80−60−40−20 0 ϖt [deg]Time (Julian centuries from 2000.0)−4−3−2−1Figure 7: Short-term evolution of the Earth's climatic precession over 13 kyr (one datum
per 120 d). The last 500 yr, from 1950, is illustrated in the insert with expanded details.
70
−0.025−0.02−0.015−0.01−0.005 0 0.005 0.01 0.015 0.02−120−100−80−60−40−20 0e sin(ϖt)Time (Julian centuries from 2000.0)−4−3−2−1Figure 8: Short-term evolution of the Earth's obliquity over 13 kyr (one datum per 120
d). The last 500 yr, from 1950, is illustrated in the insert with expanded details.
71
23.3 23.4 23.5 23.6 23.7 23.8 23.9 24 24.1 24.2 24.3−120−100−80−60−40−20 0 εt [deg]Time (Julian centuries from 2000.0)−4−3−2−1Figure 9: MTM power spectrum of the raw data from the simulated perturbations on
Earth's eccentricity by Venus to Saturn over 1 kyr (from 1000 to 2000 AD). The labels
mark the main recognized perturbations. P, P2 and P3 refer to perturbations peaks with
confidence level less than 95% which are related to planet Mars. One un-labeled peak
between P2 and 2V-3E seems to be associated with the 4E-7Ma inequality. The other
un-labeled but significant peaks are harmonics of labeled perturbations.
72
10−1510−1410−1310−1210−1110−1010−910−810−7 2 4 8 16 32V−2E2V−3E5V−8EJ3V−5EE−2MaSPP3P2Raw power spectra (V,E,Ma,J,S)period [yr]MTM of excentricityMTM 99% conf.MTM 95% conf.Figure 10: MTM power spectrum for the short-term eccentricity variations based on the
detrended data series after removing a parabolic long-term trend. The basic periodicities
identified in Fig. 9 can also be detected here.
73
10−1310−1210−1110−1010−910−810−710−610−510−410−3 2 4 8 16 32V−2E2V−3E5V−8EJ3V−5EE−2Ma 4E−7MaP2SPP32S2J3UP4P5Eccentricity − Power spectra (detrended data)Period [yr]MTMMTM 99% conf.MTM 95% conf.Figure 11: MTM power spectrum of the detrended obliquity time series. The main
periodicities are identified (see discussion in the main text). Other periodicities are from
harmonics and non-linear mixing of frequencies.
74
10−1410−1210−1010−810−610−410−2 2 4 8 16 32V−2E5V−8EJ3V−5E2S18.6 = MΩ3V−5E2J2V−3E2MΩS2.47 = 2E−3Ma4E−7Ma3E−5Ma8.85Obliquity − Power spectra (detrended data)Period [yr]MTMMTM 99% conf.MTM 95% conf.Figure 12: Daily mean irradiation at 65◦N over the indicated calendar interval for
equinoxes; i.e., λ(cid:12) t = 0, 180 deg. The solar constant used was 1366 W m−2.
75
182183184185186800100012001400160018002000YearsMarch182183184185186W [W m-2]Mean daily irradiation at 65N - EquinoxesSeptemberFigure 13: The same as Fig. 12 but for solstices; i.e., λ(cid:12) t = 90, 270 deg. The solar
constant used was 1366 W m−2.
76
477478479480800100012001400160018002000YearsJune234W [W m-2]Mean daily irradiation at 65N - SolsticesDecemberFigure 14: W at July mid-month; i.e., λ(cid:12) t = 120 deg., for 65◦N. The time interval corre-
sponds closest to the one considered in Loutre et al. (1992)'s Fig. 12a. This comparison
shows a very good agreement with their results. Solar constant used was 1366 W m−2.
77
427 427.5 428 428.5 1500 1600 1700 1800 1900TSI0 = 1366.0 W m-2Mean daily irradiance at 65N [W m-2]YearsJuly - 120 deg.Figure 15: MTM spectrum of W (over detrended time series) at July mid-month, 65◦N.
The main periodicities are identified and are described in the main text. At lower latitudes,
the periodicities arising from the obliquity signals are weakened.
78
10−810−710−610−510−410−310−210−1100101 2 4 8 16 3229.3918.6315.7611.868.097.885.925.263.983.5632.692.42W 65N, July − MTM power spectraPeriod [yr]MTMMTM 99% conf.MTM 95% conf.Figure 16: Relative difference in W calculation as a continuous λ(cid:12) t or using tabulated
values at mid-day, for solstices (λ(cid:12) t = 90, 270 deg.). At these moments, when the Sun is
stationary, the errors are small, but they can reach 0.2%.
79
-0.015-0.01-0.00500.0050.010.015-10000-8000-6000-4000-200002000YearsJune-0.015-0.01-0.00500.0050.010.015(W-W0)/W0Solstices - 65NDecemberFigure 17: The same as Fig. 16 but for equinoxes (λ(cid:12) t = 0, 180 deg.). The error is larger
on average by one order of magnitude (or about 1.2 %) when compared to the errors
during solstices.
80
-0.015-0.01-0.00500.0050.010.015-10000-8000-6000-4000-200002000YearsMarch-0.015-0.01-0.00500.0050.010.015(W-W0)/W0Equinoxes - 65NSeptemberFigure 18: Relative difference in the W calculation as a function of λ(cid:12) t over the 13 kyr
interval we studied (for 65◦N). Around the solstice of December, the relative error can
reach 5% but drop to the level of less than 0.1% at 270 deg.
81
-0.04-0.02 0 0.02 0.04 0 50 100 150 200 250 300 350 400(W-W0)/W0Solar true longitude [deg.]Figure 19: W calculation as a continuous λ(cid:12) t (lines) or using tabulated values at mid-day
(dots), for λ(cid:12) t= 250 deg. (for 65◦N), from 1000 to 2000 AD. The absolute difference is
less than 0.3 W m−2.
82
7.5 7.6 7.7 7.8 7.9 8 8.1 8.2 8.3 8.4 8.5 8.6 1000 1200 1400 1600 1800 2000Mean daily irradiance at 65N [W m-2]YearsSolar true longitue = 250 degWW0Figure 20: Difference W − W0 in mean daily insolation calculation as a function of λ(cid:12) t for
the 13 kyr interval we studied (for 65◦N). At boreal spring and at the end of the boreal
summer, the absolute difference can reach 2.5 W m−2 (see the text for more explanation).
83
-3-2-1 0 1 2 3 0 50 100 150 200 250 300 350 400(W-W0) [W m-2]Solar true longitude [deg.]Figure 21: W at July mid-month (λ(cid:12) t = 120 deg.), for 65◦N: this work, DE431-IAU
(dashed lines); Laskar et al. (2011a) solution, La2010 (points). The comparison is shown
starting at −10 kyr with expanded details of the intercomparison illustrated in the insert.
The solar constant used was 1366.0 W m−2.
84
425 430 435 440 445 450 455 460 465 470-12-10-8-6-4-2 0La2010DE431-IAUJuly Mid-month Insolation [W m-2]Milennia from J2000.0 468.85 468.9 468.95 469 469.05 469.1 469.15-10-9.8-9.6Figure 22: W at July mid-month (λ(cid:12) t = 120 deg.), for 65 ◦N: DE431-IAU solution,
continuous values (dashed lines); La2010 data (points) and DE431-IAU solution, with
tabulated values (dots). Differences between the continuous and tabulated values can
reach ± 0.3%; i.e., approximately ± 1.2 W m−2.
85
425 426 427 428 429 430 431 432-1-0.8-0.6-0.4-0.2 0July Mid-month Insolation [W m-2]Milennia from J2000.0DE431-IAU tabDE431-IAU contLa2010 |
1701.01303 | 1 | 1701 | 2017-01-05T13:01:38 | The Spitzer search for the transits of HARPS low-mass planets - II. Null results for 19 planets | [
"astro-ph.EP",
"astro-ph.IM"
] | Short-period super-Earths and Neptunes are now known to be very frequent around solar-type stars. Improving our understanding of these mysterious planets requires the detection of a significant sample of objects suitable for detailed characterization. Searching for the transits of the low-mass planets detected by Doppler surveys is a straightforward way to achieve this goal. Indeed, Doppler surveys target the most nearby main-sequence stars, they regularly detect close-in low-mass planets with significant transit probability, and their radial velocity data constrain strongly the ephemeris of possible transits. In this context, we initiated in 2010 an ambitious Spitzer multi-Cycle transit search project that targeted 25 low-mass planets detected by radial velocity, focusing mainly on the shortest-period planets detected by the HARPS spectrograph. We report here null results for 19 targets of the project. For 16 planets out of 19, a transiting configuration is strongly disfavored or firmly rejected by our data for most planetary compositions. We derive a posterior probability of 83% that none of the probed 19 planets transits (for a prior probability of 22%), which still leaves a significant probability of 17% that at least one of them does transit. Globally, our Spitzer project revealed or confirmed transits for three of its 25 targeted planets, and discarded or disfavored the transiting nature of 20 of them. Our light curves demonstrate for Warm Spitzer excellent photometric precisions: for 14 targets out of 19, we were able to reach standard deviations that were better than 50ppm per 30 min intervals. Combined with its Earth-trailing orbit, which makes it capable of pointing any star in the sky and to monitor it continuously for days, this work confirms Spitzer as an optimal instrument to detect sub-mmag-deep transits on the bright nearby stars targeted by Doppler surveys. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. aa29270
August 10, 2018
c(cid:13) ESO 2018
7
1
0
2
n
a
J
5
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
3
0
3
1
0
.
1
0
7
1
:
v
i
X
r
a
The Spitzer search for the transits of HARPS low-mass planets -
II. Null results for 19 planets⋆
M. Gillon1, B.-O. Demory2,3, C. Lovis4, D. Deming5, D. Ehrenreich4, G. Lo Curto6, M. Mayor4, F. Pepe4,
D. Queloz3,4, S. Seager7, D. S´egransan4, S. Udry4
1 Space sciences, Technologies and Astrophysics Research (STAR) Institute, Universit´e de Li`ege, All´ee du 6 Aout 17,
Bat. B5C, 4000 Li`ege, Belgium
2 University of Bern, Center for Space and Habitability, Sidlerstrasse 5, CH-3012, Bern, Switzerland
3 Cavendish Laboratory, J. J. Thomson Avenue, Cambridge CB3 0HE, UK
4 Observatoire de Gen`eve, Universit´e de Gen`eve, 51 Chemin des Maillettes, 1290 Sauverny, Switzerland
5 Department of Astronomy, University of Maryland, College Park, MD 20742-2421, USA
6 European Southern Observatory, Karl-Schwarzschild-Str. 2, D-85478 Garching bei Munchen, Germany
7 Department of Earth, Atmospheric and Planetary Sciences, Department of Physics, Massachusetts Institute of
Technology, 77 Massachusetts Ave., Cambridge, MA 02139, USA
Received date / accepted date
ABSTRACT
Short-period super-Earths and Neptunes are now known to be very frequent around solar-type stars. Improving our
understanding of these mysterious planets requires the detection of a significant sample of objects suitable for detailed
characterization. Searching for the transits of the low-mass planets detected by Doppler surveys is a straightforward way
to achieve this goal. Indeed, Doppler surveys target the most nearby main-sequence stars, they regularly detect close-in
low-mass planets with significant transit probability, and their radial velocity data constrain strongly the ephemeris
of possible transits. In this context, we initiated in 2010 an ambitious Spitzer multi-Cycle transit search project that
targeted 25 low-mass planets detected by radial velocity, focusing mainly on the shortest-period planets detected by the
HARPS spectrograph. We report here null results for 19 targets of the project. For 16 planets out of 19, a transiting
configuration is strongly disfavored or firmly rejected by our data for most planetary compositions. We derive a posterior
probability of 83% that none of the probed 19 planets transits (for a prior probability of 22%), which still leaves a
significant probability of 17% that at least one of them does transit. Globally, our Spitzer project revealed or confirmed
transits for three of its 25 targeted planets, and discarded or disfavored the transiting nature of 20 of them. Our light
curves demonstrate for Warm Spitzer excellent photometric precisions: for 14 targets out of 19, we were able to reach
standard deviations that were better than 50ppm per 30 min intervals. Combined with its Earth-trailing orbit, which
makes it capable of pointing any star in the sky and to monitor it continuously for days, this work confirms Spitzer as
an optimal instrument to detect sub-mmag-deep transits on the bright nearby stars targeted by Doppler surveys.
Key words. binaries: eclipsing -- planetary systems -- stars: individual: BD-061339, HD 1461, HD 10180, HD 13808,
HD 20003, HD 20781, HD 31527, HD 39194, HD 45184, HD 47186, HD 51608, HD 93385, HD 96700, HD 115617,
HD 125612, HD 134060, HD 181433, HD 215497, HD 219828 - techniques: radial velocity - techniques: photometric
1. Introduction
Starting from 2004 (Butler et al. 2004, Santos et al. 2004),
exoplanet search projects have been detecting planets of
a few to ∼20 Earth masses at an ever-increasing rate, re-
vealing them to be very frequent around solar-like stars
(e.g. Howard et al. 2012) where they tend to form compact
multiple systems (Rowe et al. 2014). Based on their mass
(or minimal mass for planets detected by radial velocity
- RV), these objects are loosely classified as super-Earths
(Mp ≤ 10M⊕) and Neptunes (Mp > 10M⊕). This divi-
sion is based on the theoretical limit for the runaway ac-
cretion of H/He by a protoplanet, ∼ 10M⊕ (Rafikov 2006),
and thus implicitly assumes that Neptunes are predomi-
nantly ice giants and that most super-Earths are massive
terrestrial planets. However, the growing sample of transit-
ing low-mass exoplanets with precise mass and radius mea-
surements exhibit a wide diversity of densities that reveals
a very heterogeneous population, making simplistic infer-
ences hazardous when based on the comparison with solar
system planets. A better understanding of this ubiquitous
class of planets requires the thorough characterization of
a significant sample, not only the precise measurements of
their physical dimensions but also the exploration of their
atmospheric composition to alleviate the strong degenera-
cies of composition models in this mass range (e.g. Seager
et al. 2007, Valencia et al. 2013).
Send offprint requests to: [email protected]
⋆ The photometric and radial velocity time series used in
this work are only available in electronic form at the CDS
via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via
http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/
Among the known transiting low-mass planets, GJ 436 b
(Butler et al. 2004, Gillon et al. 2007) and GJ 1214 b
(Charbonneau et al. 2009) are the most thoroughly char-
acterized planets in the Neptune and super-Earth mass
1
M. Gillon et al.: Results of the Warm Spitzer transit search
ranges, respectively, thanks to the small size (∼ 0.45 and 0.2
R⊙) and proximity (about a dozen of pc) of their M-dwarf
host stars. First constraints on their atmospheric proper-
ties have indeed been obtained by several programs (e.g.
Kreidberg et al. 2014, Knutson et al. 2014a, Ehrenreich
et al. 2015). Extending this kind of detailed studies to
Neptunes and super-Earths orbiting solar-type hosts re-
quires the detection of such planets in transit in front of ex-
tremely bright and nearby stars. A straightforward method
to achieve this goal is to search for the transits of the low-
mass planets detected by RV surveys. Indeed, these sur-
veys target the most nearby main-sequence stars, and they
have now detected enough short-period low-mass planets
to make it highly probable that a handful of them transit
their parent stars, as demonstrated by the previous detec-
tions by the MOST space telescope of the transits of the
super-Earths 55 Cnc e (Winn et al. 2011)1. Thanks to the
brightness of their host stars, the atmospheric characteriza-
tion of these two planets has already started (e.g. Demory
et al. 2012, Knutson et al. 2014b). More recently, the same
approach enabled us to reveal with Spitzer the transiting
configuration of the rocky planet HD 219134 b (Motalebi et
al. 2015) which, at 6.5pc, is the nearest known transiting
exoplanet.
Searching for the transits of RV low-mass planets is one
of the main objectives of the future European space mission
CHEOPS (Broeg et al. 2013). However, CHEOPS is not due
to launch before the end of 2017. To set it on its path, back
in 2010 we set up an ambitious project using Spitzer/IRAC
(Fazio et al. 2004) to search for the transits of the RV low-
mass planets that have the highest geometric transit proba-
bilities, focusing mainly on the shortest-period planets de-
tected by the HARPS spectrograph (Mayor et al. 2003).
Our Spitzer transit search was composed of a cryogenic
program targeting HD 40307 b (ID 495, 27.5hr), and three
so-called Warm (i.e. non cryogenic) programs (ID 60027,
90072, and 11180; 100hr, 300hr, and 9.5hr) that targeted
24 other RV low-mass planets. Its published results have
so far been the non-detection of the transits of HD 40307 b
(Gillon et al. 2010, hereafter G10) and GJ 3634 b (Bonfils
et al. 2011), the detection and confirmation of the tran-
sits of 55 Cnc e (Demory et al. 2011, Gillon et al. 2012a),
the confirmation of the transiting nature of HD 97658 b
(Van Grootel et al. 2014), and the detection of a transit
of HD 219134 b (Motalebi et al. 2015). Results for another
planet will be presented in a forthcoming paper (S´egransan
et al., in prep.) We report here null results for the 19 other
targets of the project.
We first present our targets and our determination of
their transit ephemeris. In Sect. 3, we present our Spitzer
data and their reduction. Section 4 describes our data anal-
ysis. Its main results are presented in Sect. 5. Finally, we
discuss our global results and give our conclusions in Sect.
6.
2. Targets and transit ephemeris determination
Table 1 and Tables A.1 to A.4 list the 19 targets of this
work. For each target, we performed an analysis of the avail-
able RVs, including new measurements for some of them
1 Our Spitzer program independently revealed the transit-
ing nature of 55 Cnc e (Demory et al. 2011). and HD 97658 b
(Dragomir et al. 2013)
2
gathered by HARPS2, to derive the most accurate tran-
sit ephemeris. This analysis was done with the adaptative
Markov Chain Monte-Carlo (MCMC) algorithm described
in G10 (see also Gillon et al. 2012a, 2012b, 2014). Our nom-
inal model was based on a star and one or several planets
on Keplerian orbits around their common center of mass.
For some cases, we added a linear or quadratic trend to
the model, based on the minimization of the Bayesian in-
formation criterion (BIC, Schwarz 1974) to elect our final
model. We checked that planet-planet interactions had neg-
ligible influence on our solutions, using, for this purpose,
the Systemic Console software (Meschiari et al. 2009). We
also used this software to perform an initial optimization of
parameters to check that no solution existed with a higher
likelihood than the published one. During this initial stage
of the analysis, the RV errors were assumed to be measure-
ment uncertainties. Once this initial stage was completed,
we measured the quadrature difference between the rms of
the residuals and the mean error, and this "jitter noise"
(Wright 2005) was quadratically summed to the measure-
ment uncertainties. We then performed a MCMC analy-
sis to probe the posterior probability distribution functions
(PDF) of the model and physical parameters
Each MCMC analysis was composed of five Markov
chains of 105 steps, the first 20% of each being considered
as its burn-in phase and discarded. For each run, the con-
vergence of the five Markov chains was checked using the
statistical test of Gelman & Rubin (1992). The resulting
posterior PDFs for the transit time and period were then
used to schedule the Spitzer observations, with monitoring
the 2 − σ transit window as the goal, i.e. keeping the prob-
ability to miss a transit below 5%.
Even if the resulting posterior PDFs for the orbital
eccentricity of most target planets were compatible with
zero, we did not assume a purely circular orbit for them
to ensure the reliability of our derived transit ephemeris.
Indeed, several examples of low-mass exoplanets with sig-
nificantly eccentric few-days orbits (e.g. GJ 436 b, HAT-P-
11 b, HD215497 b) remind us of the risks of systematically
assuming circular orbits for close-in exoplanets based on
tidal circularization arguments. An established theory for
tidal dissipation mechanisms is still a long-term goal, and
reaching it relies mostly on gathering new observational
constraints.
Once the Spitzer observations had been performed for a
planet, we made a second MCMC analysis of the most up-
to-date RV dataset, this time with the fitted transit time
corresponding to the epoch covered by the Spitzer run. The
purpose of this analysis was to derive the most accurate or-
bital parameters, transit ephemeris, and minimal masses for
the planets under consideration. Indeed, for most of them
HARPS gathered a significant number of RV measurements
between the scheduling of the Spitzer observations and the
final analysis of the Spitzer images.
Tables A.1 to A.4 present the results obtained for each
targeted planet from these MCMC analyses of the most up-
to-date RV dataset. In addition to some basic parameters
for the host stars, these tables give the origin of the RVs
used as input in our MCMC analysis for each target, and
provide the most relevant results of our MCMC analysis:
2 Most HARPS measurements
in
are
on the ESO/HARPS online
http://archive.eso.org/wdb/wdb/eso/repro/form
available
used
this work
at
archive
M. Gillon et al.: Results of the Warm Spitzer transit search
the transit and occultation ephemeris, the minimal mass,
the expected minimum transit depth corresponding to a
pure iron planet (Seager et al. 2007), the equilibrium day-
side temperature, the orbital parameters, and the expected
duration for a central transit. The radius of the star was de-
rived from the luminosity and effective temperature, taking
bolometric corrections from Flower (1996).
Tables A.1 to A.4 also present the median value and the
1-σ errors for the prior transit probability. It was computed
at each step of the MCMC with the following formula:
P (tr) = (cid:18) R∗
a (cid:19)(cid:18) 1 + e sin ω
1 − e2 (cid:19),
(1)
where R∗ is the stellar radius, a is the orbital semi-major
axis, e is the orbital eccentricity and ω is the argument
of periastron. This probability estimate does not take into
account that planets are more likely to be discovered by
RV if their orbit is significantly inclined (e.g. Wisniewski
et al. 2012). By performing Bayesian simulations that as-
sume different prior PDFs for the planetary masses, Stevens
& Gaudi (2013) have shown that this bias increases the
transit probability of short-period low-mass RV planets
like the ones considered here by only ∼20% in average.
Furthermore, its actual estimation for a given planet de-
pends strongly on the assumed prior PDFs for the plane-
tary masses. We have thus neglected it in the context of
this work.
target in Tables A.5 to A.11. These tables also give the ver-
sion of the Spitzer pipeline used to calibrate the correspond-
ing images, the resulting files being called basic calibrated
data (BCD) in the Spitzer nomenclature. Each subarray
mode BCD is composed of a cube of 64 subarray images of
32×32 pixels (pixel scale = 1.2 arc second).
The following reduction strategy was used for all the
Spitzer AOR. We first converted fluxes from the Spitzer
units of specific intensity (MJy/sr) to photon counts, then
aperture photometry was performed on each subarray im-
age with the IRAF/DAOPHOT5 software (Stetson, 1987). For
each AOR, we tested different aperture radii and back-
ground annuli, and selected the combination minimizing
the white and red noises in the residuals of a short data
fitting analysis. The center and width of the point-spread
functions (PSF) were measured by fitting a 2D-Gaussian
profile on each image. The x − y distribution of the mea-
surements was then studied, and measurements that had a
visually discrepant position relative to the bulk of the data
were then discarded. For each block of 64 subarray images,
we then discarded the discrepant values for the measure-
ments of flux, background, x- and y-positions using a 10-σ
median clipping for the four parameters, and the resulting
values were averaged, the photometric errors being taken as
the errors on the average flux measurements. Finally, a 50-
σ slipping median clipping was used on the resulting light
curves to discard outliers (owing to, for example, cosmic
hits).
3. Warm Spitzer photometry
Tables A.5 to A.11 provide a summary of the Spitzer ob-
servations. Since all our targets are very bright (K between
2.96 and 6.90), all of them were observed in subarray mode
(32x32 pixels windowing of the detector), the extremely fast
Fowler sampling (∼0.01s), which maximizes the duty cycle
and signal-to-noise ratio (S/N). No dithering pattern was
applied to the telescope (continuous staring). For each tar-
get, the exposure time was selected to maximize the S/N
while staying in the linear regime of the detector, basing on
the Warm Spitzer flux density estimator tool3 and on the
instructions of the Warm Spitzer Observer Manual4.
The observations of program 60027 (Cycle 6) and 90072
(Cycle 9) were performed between 2009 Dec 14 and 2010
Sep 11, and between 2012 Dec 03 and 2014 May 14, re-
spectively. For several of the Cycle 9 targets, we benefitted
from the newly introduced PCRS peak-up mode (Ingalls
et al. 2012). This mode provides enhanced accuracy in the
position of the target on the detector, to a significant de-
crease of the so-called 'pixel phase effect' that is the most
important source of correlated noise in high-S/N staring
mode observation with IRAC InSb arrays (e.g. Knutson et
al. 2008). For HD 1461 b, we supplemented our data with
the IRAC 4.5µm observations presented by Kammer et al.
(2014), as described in Sect. 5.2.
On a practical level, each observation run was divided in
one or several science astronomical observational requests
(AOR) of 12hr at most, preceded by a short (20-30 min)
AOR to allow the pointing of the telescope and the instru-
ment to stabilize. The IDs of all AORs are given for each
3 http://ssc.spitzer.caltech.edu/warmmission/propkit/pet/starpet
4 http://ssc.spitzer.caltech.edu/warmmission/propkit/som
4. Global Warm Spitzer + RV data analysis
We analyzed the Spitzer photometric time-series supple-
mented by the RVs with our MCMC code. For each target,
our model for the RVs was the same as the one presented
in Sect. 2. The assumed photometric model consisted of
the eclipse model of Mandel & Agol (2002) to represent
the possible eclipses of the probed planets, multiplied for
each light curve by a baseline model that aimed to repre-
sent the other astrophysical and instrumental effects at the
source of photometric variations. We assumed a quadratic
limb-darkening law for the stars. For each light curve that
corresponded to a specific AOR, we based the selection of
the baseline model on the minimization of the BIC. Tables
A.5 to A.11 present the baseline function elected for each
AOR.
Following Gillon et al. (2014), the instrumental mod-
els included three types of low-order polynomials. The first
one had as variables the x- and y-positions of the center of
the PSF to represent the so-called pixel phase effect of the
IRAC InSb arrays (e.g. Knutson et al. 2008). The second
one had the PSF widths in the x- and/or the y-direction
as variables, its inclusion in the baseline model strongly in-
creasing the quality of the fit for all AORs (the so-called
PSF breathing effect, see also Lanotte et al. 2014). The
third function was a polynomial of the logarithm of time to
represent a sharp increase of the detector response at the
start of some AORs (the so-called ramp effect, Knutson et
al. 2008). To improve the quality of the modeling of the
pixel phase effect, especially the fitting of its highest fre-
5 IRAF is distributed by the National Optical Astronomy
Observatory, which is operated by the Association of
Universities for Research in Astronomy, Inc., under cooperative
agreement with the National Science Foundation.
3
M. Gillon et al.: Results of the Warm Spitzer transit search
quency components,for most AORs we supplemented the
x- and y-polynomial with the bi-linearly-interpolated sub-
pixel sensitivity (BLISS) mapping method (Stevenson et al.
2012). The sampling of the position space was selected so
that at least five measurements fall within the same sub-
pixel. See Gillon et al. (2014) for more details.
The jump parameters of the MCMC, i.e. the parameters
randomly perturbed at each step of the Markov Chains,
were as follows:
-- The stellar mass M∗, radius R∗, effective temperature
Tef f , and metallicity [Fe/H]. For these four parameters,
normal prior PDFs were assumed based on the values
given in Tables A.1-4.
-- For the potential transiting planet, the parameter b′ =
a cos i/R∗, where a is the orbital semi-major axis and
i is the orbital inclination. b′ would correspond to the
transit impact parameter in the case of a circular orbit.
The step was rejected if b′ > a/R∗. For the other planets
of the system, b′ was fixed to 0.
-- The parameter K2 = K√1 − e2 P 1/3 for all planets of
the system, K being the RV orbital semi-amplitude, e
the orbital eccentricity, and P the orbital period.
-- The orbital period P of each planet.
-- For each planet, the two parameters √e cos ω and
√e sin ω, e being the orbital eccentricity and ω being
the argument of periastron.
-- The planet/star area ratio dF = (Rp/R∗)2 for the po-
tential transiting planet. At each step of the MCMC, the
planetary radius corresponding to a pure iron composi-
tion was computed under the formalism given by Seager
et al. (2007), and if the planetary radius derived from
dF and R∗ was smaller, the step was rejected. A sim-
ilar rejection was done for Rp > 11R⊕, an implausibly
large size for the low-mass planets considered here. The
goal of these prior constraints on Rp was to avoid fit-
ting extremely shallow transits and ultra-grazing tran-
sits of unrealistically big planets in the correlated noise
of the light curves to ensure an unbiased posterior tran-
sit probability. For the other planets of the multiple sys-
tems, dF was fixed to 0. In all cases, we checked that
a transit of another planet of these systems was not
expected to occur during the Spitzer observations.
-- The time of inferior conjunction T0 for all planets of
the system. For the potential transiting planets that we
considered, T0 corresponds approximatively to the mid-
time of the transit searched for by Spitzer.
The limb-darkening of the star was modeled by a
quadratic law (Claret 2000). For both Warm Spitzer band-
passes (3.6 and 4.5 µm), values for the two quadratic limb-
darkening coefficients u1 and u2 were drawn at each step
of the MCMC from normal distributions whose expecta-
tions and standard deviations were drawn from the tables
of Claret & Bloemen (2011) for the corresponding band-
passes and for the stellar atmospheric parameters given in
Tables A.1-4.
Five chains of 100000 steps were performed for each
analysis, their convergences being checked using the sta-
tistical test of Gelman and Rubin (1992). They followed
a preliminary chain of 100000 steps, which was performed
to estimate the need to rescale the photometric errors. For
each light curve, the standard deviation of the residuals was
then compared to the mean photometric errors, and the re-
sulting ratios βw were stored. βw represents the under- or
4
overestimation of the white noise of each measurement. On
its side, the red noise present in the light curve (i.e. the
inability of our model to represent perfectly the data) was
taken into account as described in G10, i.e. a scaling fac-
tor βr was determined from the standard deviations of the
binned and unbinned residuals for different binning inter-
vals ranging from 5 to 120 minutes, the largest values being
kept as βr. At the end, the error bars were multiplied by
the correction factor CF = βr × βw. The derived values for
βr and βw are given for each light curves in Tables A.5-11.
5. Results
For each planet searched for transit, Table 1 presents the
derived posterior full transit probability P (f, D), i.e. the
probability that the planet undergoes full transits given the
Spitzer data. Bayes theorem shows that
P (f, D) =
P (f )P (D, f )
P (f )P (D, f ) + P (g)P (D, g) + P (n)P (D, n)
(2)
P (f ), P (g), and P (n) are the prior (geometric) probabil-
ities of full, grazing and no transit, respectively (P (g) is
close to zero, and P (n) ∼ 1 − P (t)). P (D, f ), P (D, g), and
P (D, n) are the probabilities (likelihoods) to have the ob-
served data given the three mutually exclusive hypotheses.
All the terms of the right-hand side of Eq. 2 are probed
by the MCMC analysis, resulting in accurate estimates of
P (f, D).
Figs. 1-4, 6-14, and 16-20 show the resulting detrended
Spitzer light curves and the derived posterior PDFs for the
inferior conjunction. Below, we provide relevant details for
our 19 targets.
5.1. BD-061339
BD-061339 (a.k.a. GJ 221) is a V=9.7 late-K dwarf around
which two planets were detected by HARPS (Lo Curto et
al. 2013), a super-Earth on a ∼3.9d orbit and a planet of ∼
50M⊕ minimal mass on a ∼126d orbit. Our model selection
process for the RVs (HARPS + PFS) favored a model with
a slope in addition to these two planets (Table A.1) . The
derived value for this slope is −0.73 ± 0.15 m s−1 per year,
which could correspond to a giant planet in outer orbit or
to the imprint of a stellar magnetic cycle.
As can be seen in Fig. 1, the high-precision of the Spitzer
photometry (42 ppm per 30 min time bin) discards any
transit for BD-061339 b during our observations, even for
an unrealistic pure iron composition. However, the edge of
the right wing of the PDF for T0 was not explored by our
observations, so a late transit is still possible. The resulting
posterior full transit probability is 0.53% (Table 1), high
enough to justify a future exploration of the late part of
the transit window.
5.2. HD 1461
HD 1461 is a V=6.5 solar-type star known to host a super-
Earth on a 5.77 d orbit (Rivera et al. 2010). Recently, a
second super-Earth on a 13.5 d orbit was announced by
D´ıaz et al. (2015). In our analysis of the HARPS + Keck
RVs for this star, the convergence of the MCMC appeared
to be significantly improved by removing the 20 first Keck
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 1. Warm Spitzer photometry for BD-061339 after cor-
rection for the instrumental effects, normalization, and bin-
ning per 30 min. Models for a central transit of BD-061339 b
are shown, assuming a pure iron (red solid line) and pure
water ice (blue dashed line) composition. The highest pos-
terior probability model, which assumes no transit, is also
shown (purple line). The posterior PDF for the transit tim-
ing derived from the RV analysis is shown above the figure.
measurements for which the exposure time and the result-
ing precision were significantly lower than for the rest of
the data. Our analysis favored a three planet model (5.77,
13.5, and 377 days period) in addition to a second-order
time polynomial, suggesting the presence of a fourth low-
frequency signal, and an activity model consisting of a sum
of second-order polynomials in the cross-correlation func-
tion (CCF, Baranne et al. 1996, Queloz et al. 2001) parame-
ters (Table A.1): width, contrast, and bisector. We inferred
that the 377d period, close to the duration of a year, is
caused by a systematic effect (the stitching) recently re-
vealed to affect HARPS data (Dumusque et al. 2015). The
data used by D´ıaz et al. (2015) were corrected for this ef-
fect. We made several tests that showed that the inclusion
or not of this 377d Doppler signal in the RV model did
not affect the results for the closest-in planet, including its
transit ephemeris, so we kept it in our final analysis.
A search for a transit of HD 1461 b with Spitzer was pre-
sented by Kammer et al. in 2014 (program 80220). We first
reduced their data and used them with the HARPS+Keck
RVs as input for a global MCMC analysis. The resulting
posterior full transit probability was 0.5%, suggesting that
a small but significant fraction of the transit window was
not covered by these Spitzer observations. As can be seen in
Fig. 2, a transit that had ended just before the Spitzer ob-
servations remained possible. This possibility is amplified
by the ramp effect that affected the first hour of the Spitzer
data, resulting in an increase of brightness that could be
degenerated with a transit egress. We thus complemented
these Spitzer archive data with new Spitzer observations
covering the first part of the transit window. We performed
a global analysis of all Spitzer + RVs that made a transit-
ing configuration for HD 1461 b very unlikely (see Fig. 2),
the resulting posterior full transit probability now being of
only 0.14% (Table 1).
Fig. 2. Same as Fig. 1 for HD 1461 b. Upper panel cor-
responds to archive data (program 80220, Kammer et al.
2014), and the lower panel corresponds to the new data
gathered in our program 90072. The upper panel also shows
the light curve derived without correction of the ramp effect
(open pale green circles).
5.3. HD 10180
HD 10180 is a V=7.3 solar-type star hosting a particularly
interesting system of six planets of relatively low-masses
(Lovis et al. 2011). At first, we did not consider search-
ing for the transit of the 1.18d-period Earth-mass planet
candidate HD 10180 b presented by Lovis et al. (2011) be-
cause its Doppler detection was not firmly secured when
we planned our Spitzer Cycle 6 observations of the star.
Furthermore, we estimated that its transit would, in any
case, be too shallow (a few dozens of ppm, at most 100
ppm) to be firmly detected with Spitzer. We thus focused
on the planet HD 10180 c (M sin i = 13M⊕, P = 5.76 d),
for which the Doppler signal was clearly detected in the
HARPS data, and for which the expected transit depth
(> 150 ppm) was large enough to ensure a sure detection
with Spitzer. enabled us to discard a transit of the planet
(Fig. 3), the resulting posterior full transit probability be-
ing of 0.14% (Table 1). Still, we noticed a shallow structure
in the detrended photometry that occurs at a time con-
sistent with a transit of HD 10180 b (Fig. 4), the best-fit
transit depth of ∼90 ppm translating into a radius of 1.2
R⊕ consistent with a planetary mass of 1M⊕. Based on
this result, we decided to observe in our Cycle 9 program
90072 two more transit windows of HD 10180 b, this time
at 4.5 µm. No transit structure was clearly detected in the
resulting light curves (Fig. 4, bottom), and a short MCMC
analysis of the data led us to conclude that these new data
do not increase the significance of the 2010 tentative detec-
tion. We thus conclude that the low-amplitude structure in
5
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 3. Same as Fig. 1 for the light curve of 2010 Jan 16
for HD 10180 c.
our Cycle 6 Spitzer light curve is probably just correlated
noise of instrumental origin that is not perfectly represented
by our baseline model and that can easily be modeled by
a shallow enough transit profile. For Spitzer, the average
amplitude of these correlated noise structures is of a few
dozen of ppm (see Sec. 6), making the firm detection of a
unique transit shallower than ∼100ppm impossible - simi-
larly to the one expected for HD 10180 b (see discussion in
Sec. 6 and our estimated detection thresholds in Table 1).
This so-called red noise limit can be surpassed, but only by
gathering more observations of the transit window, as we
did here (see also the case of HD 40307 b in Gillon et al.
2010).
5.4. HD 13808
In 2012, we analyzed the HARPS dataset for the V =8.4
early K-dwarf HD 13808, confirming the existence of
the two planets around it with periods of 14.2 d and
53.7 d announced by Mayor et al. (2011, hereafter M11).
Furthermore, our analysis revealed the existence of (1)
a low-frequency signal that was well-modeled with a
quadratic trend, whose origins is the magnetic cycle of
the star (Queloz et al. in prep.), and (2) a low-amplitude
Doppler signal with a period of 1.091 d which corresponds
to a planet of Mp sin i = 1.5± 0.3 M⊕ with an interestingly
high transit probability of ∼20%. The false-alarm probabil-
ity (FAP) derived by Systemic for this short-period planet
candidate was close to 1%. Based on this small FAP and the
scientific importance of this putative planet if transiting, we
decided to monitor two of its transit windows with Spitzer.
Unfortunately, the resulting light curves did not show any
convincing transit-like structure (Fig. 5). Our Spitzer data
were not acquired during a transit window of the planet at
14.2 d, thus keeping its transiting nature unconstrained.
In 2014, we analyzed the updated HARPS dataset. This
analysis confirmed the existence of the low-frequency signal
of magnetic cycle origins, still well-modeled by a quadratic
trend, but it led to a much lower significance (FAP ∼ 10%)
for the 1.091 d signal. A stronger signal at 18.98 d and at
its alias period of 1.056 d emerged with FAP close to 1%.
Still, including a polynomial function of the CCF param-
eters (bisector, contrast, width) in our MCMC modeling,
6
Fig. 4. T op: Same as Fig. 3, except that a possible tran-
sit of HD 10180 b was included in the global model. Highest
posterior probability transit model for HD 10180 b is shown
in purple. The posterior PDF for the transit timing of
HD 10180 b derived from an RV analysis assuming 7 planets
is shown above the figure. Bottom: 2013 Warm Spitzer pho-
tometry for HD 10180 binned per 30 min, corrected for the
instrumental effects, normalized, and folded on the transit
ephemeris for HD 10180 b based on our tentative detection
of a transit and on the results of an RV analysis assuming
7 planets (dT = time from most likely mid-transit time).
these signals vanished from the residuals, indicating that
their origin is stellar (activity), not planetary. The period
18.98 d could then be interpreted as the rotation period
of the star, resulting in an equatorial rotation speed ∼2.2
km.s−1 which is consistent with the measured v sin i of 2
km.s−1 (Glebocki & Gnacinski 2005). This stellar signal,
combined with aliasing effects, would thus be responsible
for creating spurious signals with periods slightly larger
than 1 d.
5.5. HD 20003
HD 20003 is a late-G dwarf of magnitude V =8.4 for which
M11 announced the detection by HARPS of two planets
with periods of 11 d and 33 d. Our analysis of the up-
dated HARPS dataset for HD 20003 not only confirmed
these planets, but also revealed two longer period signals:
one at ∼180 d that was determined as originating from the
stitching effect, and another, at ∼10 yr, that originates from
the magnetic cycle of the star (Udry et al. in prep.). Only
HD 20003 b (P =11 d) has a significantly eccentric orbit. We
used Spitzer to probe its transiting nature. We did not de-
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 5. Same as Fig. 1 for HD 13808 b. These data probed
the 2-σ transit window of a planetary candidate that turned
out to be a spurious signal with more RV data (see Sec. 5.4
for details).
Fig. 7. Same as Fig. 1 for HD 20781 b.
cially interesting target for a transit search, so we included
it in the target list of our Spitzer program and observed
one of its transit window during Cycle 9. Because of the
extreme faintness of the Doppler signal (K ∼ 90 cm.s−1),
the orbital eccentricity of the planet is poorly constrained
from the RVs alone, resulting in a particularly large transit
window. As such, we assumed the orbit to be circular to
minimize the required Spitzer time, judging a circular or-
bit as a reasonable assumption, taking into consideration
the strong tidal forces exerted by the star at such close dis-
tance and the compactness of the planetary system that
would make any significantly eccentric orbit unstable.
Our Spitzer photometry did not reveal any transit-like
structure (Fig. 7), the resulting posterior full transit proba-
bility being 0.15% (Table 1). Still, our precision is not high
enough to securely detect the transit of this very low-mass
planet if its amplitude is below ∼150ppm, corresponding
to Mercury-like iron-rich compositions. Re-observing the
planet's transit window at higher precision with, for exam-
ple, CHEOPS will be mandatory to fully exclude a transit-
ing configuration for the planet.
Fig. 6. Same as Fig. 1 for HD 20003 b.
5.7. HD 31527
tect a transit (Fig. 6). But, as can be seen in Fig. 6, we did
not probe the latest part of its transit window, resulting in
a small but still significant posterior full transit probability
of 0.54% (Table 1).
5.6. HD 20781
Based on 96 HARPS measurements, M11 announced the
discovery of two Neptunes with orbital periods of 29.1 d
and 85.1 d around this nearby V =8.4 K0-dwarf. The anal-
ysis of our much extended HARPS dataset (212 points)
confirmed the existence of these two planets, while re-
vealing the existence of two super-Earths in shorter or-
bits: a Mp sin i = 6.3M⊕ planet at 13.9 d period, and a
Mp sin i = 2.1M⊕ planet at 5.3 d period (Udry et al. in
prep.). The low minimal mass and relatively high transit
probability (>7%) of this latter planet made it an espe-
M11 announced the existence of three Neptune-mass plan-
ets around this nearby V = 7.5 solar-type star, with orbital
periods of 16.5 d, 51.3 d, and 275 d. Our analysis of the ex-
tended HARPS dataset (242 vs 167 measurements) con-
firmed the existence of these planets and improved their
orbital parameters, while not revealing any other planet
(Udry et al. in prep.). We used ∼36 hr of continuous Spitzer
observation to search for a transit of the innermost planet,
HD 31527 b (geometric transit probability = 4.4%). The re-
sulting light curve did not reveal any transit (Fig. 8), the
resulting posterior full transit probability being of 0.45%
(Table 1). Compared to the orbital solution that we used
to schedule our Spitzer observations, the updated solution
presented here (Table 2) results in Spitzer observations that
are not well centered on the peak of the posterior PDF for
the transit timing. The right wing of this PDF is thus unex-
plored. Its future exploration with, for example, CHEOPS
would be desirable to fully exclude a transiting nature for
the planet.
7
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 8. Same as Fig. 1 for HD 31527 b.
Fig. 10. Same as Fig. 1 for HD 45184 b.
Fig. 9. Same as Fig. 1 for HD 39194 b.
Fig. 11. Same as Fig. 1 for HD 47186 b.
5.8. HD 39194
This star is a V = 8.1 early K-dwarf around which HARPS
detected three super-Earths with orbital periods of 5.6 d,
14.0 d, and 33.9 d (M11, Queloz et al. in prep.). Our anal-
ysis of the updated HARPS dataset (261 RVs vs 133 in
the discovery paper) fully confirmed the existence of these
planets, and revealed a low-amplitude trend. With a geo-
metric transit probability of ∼6.5%, theinnermost of these
planets was an interesting target for our program, and we
monitored one of its transit windows with Spitzer in Dec
2012. As shown in Fig. 9, the resulting light curve was flat,
our deduced posterior transit probability being of 0.46%
(Table 1).
5.9. HD 45184
M11 reported the detection of a Neptune-mass planet on a
5.9 d period around this bright (V = 6.4) solar-type star.
Based on more than double the HARPS measurements (174
vs 82), our analysis confirms the existence of this planet,
while revealing the presence of a second planet of simi-
lar mass, ∼9.5 M⊕, on a outer orbit (P =13.1 d), and a
trend in the RVs that we could relate to the magnetic cycle
of the star (Udry et al. in prep.). We monitored the star
8
for more than 11hrs with Spitzer to search for the transit
of HD 45184 b. The resulting light curve is flat (Fig. 10),
while its precision would have been high enough to detect
the searched transit for any plausible composition. We did
not explore the latest part of the transit window, leaving a
posterior probability of 1.3% that the planet undergoes full
transit (Table 1), for a prior geometric probability of 7.7%.
5.10. HD 47186
HD 47186 b is a short-period (P =4.08 d) Neptune-mass
planet discovered by HARPS in 2009 (Bouchy et al. 2009)
around a V = 7.6 solar-type star, which is also orbited by
a giant planet at a much longer period. Our analysis of the
extended HARPS dataset confirmed the existence of these
two planets, and enabled us to derive an excellent precision
on the time of inferior conjunction of the inner planet (1σ-
error of 30 min for our Spitzer observations). We searched
for its transit within our Spitzer cycle 6 program. The re-
sulting flat light curve (Fig. 11) enables us to fully reject
the transiting nature of the planet (Table 1).
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 12. Same as Fig. 1 for HD 51608 b.
Fig. 13. Same as Fig. 1 for HD 93385 d.
5.11. HD 51608
We used Spitzer to search for the transit of the P = 14.07 d
Neptune-mass planet HD 51608 b detected by M11 around
this bright (V = 6.3) late G-type dwarf. Our analysis of the
updated HARPS dataset (210 RVs) confirmed the existence
of the two planets (P = 14.07d and 95d) announced in M11,
and favored a low-amplitude trend of probable magnetic
cycle origin (Udry et al. in prep.). Our Spitzer photometry
(Fig. 12) enabled us to discard a transit of HD 51608 for
any plausible planetary composition, our posterior full tran-
sit probability being 0.11% (Table 1) for a prior geometric
transit probability of ∼4%.
5.12. HD 93385
The detection of two low-mass planets was reported by
M11 for this V = 7.5 solar-type star, with derived mini-
mal masses of 8.4 and 10.1 M⊕ and orbital periods of 13.2 d
and 46 d, respectively. Our analysis of the updated HARPS
dataset (231 RVs) revealed the existence of a third lower-
mass (M p sin i = 4.0 ± 0.5M⊕) planet on a 7.3 d orbit
(Queloz et al. in prep., see Table A.3). With Spitzer, we
monitored a transit window (∼29hr) of this new planet,
HD 93385 d. Our resulting photometry (Fig. 13) did not
reveal any clear transit-like structure, and the posterior
full transit probability that we derived from its analysis
is 0.23% (Table 1), for a prior geometric transit probabil-
ity of 7.9%. By injecting transit models in this light curve
and analyzing the results with our MCMC code, we con-
cluded that its precision of ∼ 40 ppm per half-hour is good
enough to discard any central transit of a planet with a
density equal or smaller than Earth's, but cannot firmly
discard the transit of denser planet. We thus recommend
the re-observation of the transit window at higher photo-
metrical precision with, for example, CHEOPS.
5.13. HD 96700
HD 96700 b is a ∼ 9M⊕ planet on a 8.1 d orbit discovered by
M11 around a V = 6.5 solar-type star. Our analysis of the
updated HARPS dataset (244 RVs) confirmed its existence
and the outer Neptune-mass planet found by M11 on a
∼100 d period. It also revealed a low-amplitude trend in
Fig. 14. Same as Fig. 1 for HD 96700 b.
the RVs of probable magnetic cycle origin (Queloz et al.
in prep.). We searched for the transit of HD 96700 b with
Spitzer in 2013, without success (Fig. 14). The precision of
the Spitzer light curve is high enough to discard a transit
of HD 96700 b in it for any possible planetary composition.
The resulting posterior transit probability of the planet is
0.09%, for a prior probability of 6.8% (Table 1).
5.14. HD 115617
HD 115617 (aka 61 Vir) is a V =4.7 solar-type star (G5V) at
only 8.5 pc from Earth. A close-in super-Earth (Mp sin i =
5M⊕, P = 4.215 d) and two outer Neptunes (Mp sin i = 18
& 23M⊕, P = 38 & 123 d) were discovered around it by
Vogt et al. (2010) using RVs obtained with Keck/HIRES
and the Anglo-Australian Telescope (AAT). We performed
a global analysis of the Keck, and AAT RVs that confirmed
the existence of the three planets, without revealing any ad-
ditional object orbiting the star. We used Spitzer to observe
a transit window of HD 115617 in March 2010.
With K = 2.96, HD 115617 is an extremely bright star
for Spitzer. At 4.5 µm, it is nevertheless faint enough to
be unsaturated for the shortest available integration time
(0.01s). However, the Spitzer Science Center (SSC) in-
formed us that the requested long observation of HD 15617
9
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 15. T op: 61 Vir light curve obtained by Spitzer to
search for the transit of its planet b, divided for the best-
fit phase-pixel model (here a third order x and y-position
polynomial), unbinned (green dots) and binned to inter-
vals of 0.005d (7.2 min). The start of the second AOR is
represented by the red vertical line. Bottom: Lomb-Scargle
periodogram showing a clear power excess at ∼0.11 days
and its first harmonic.
could not be performed at 4.5 µm for technical reasons.
At 3.6 µm, the star has a flux density of 19000 ± 4000
mJy while the saturation limit is 20000 mJy for an integra-
tion time of 0.01s. SSC informed us the observations could
still be performed without any risk of saturation if the star
was placed at the corner of four pixels, leading to an effec-
tive flux density that was low enough to avoid saturation.
We decided to test this strategy. Unfortunately, the light
curve corrected for known systematic effects is corrupted
by a clear variability at the level of a few tens of percent
(Fig. 15, top). A Lomb-Scargle periodogram (Press et al.
1992) of the corrected photometry reveals a power excess
at ∼0.11 days (Fig 15, bottom). We could not identify the
origin of this variability. Nevertheless, an astrophysical ori-
gin is very unlikely, since HD 115617 is known to be an old
inactive star (Vogt et al. 2010). This variability probably
originates from the near-saturation of the detector. Indeed,
the center of the stellar image was not located at the corner
of four pixels as intended. The instrumental effect at work
has a typical timescale similar to the signal we are trying
to detect, and its amplitude is much larger than the one of
the searched transit. Even worse, it seems highly variable
in nature and we could not find any analytical function of
external parameters (PSF peak value, center position and
width, background, etc.) able to represent it satisfactorily.
Without a thorough understanding of the effect, we con-
cluded that searching for a transit of a few hundreds of
ppm with such data was illusory. The transiting nature of
HD 115617 b thus remains unconstrained by our project.
10
Fig. 16. Same as Fig. 1 for HD 125612 c.
5.15. HD 125612
HD 1215612 is a V = 8.3 solar-type star around which
three planets have been detected so far by RVs, first a P ∼
510 d gas giant in 2007 (Fischer et al.) with Keck+HIRES
data, then a close-in (P =4.15 d) Neptune-mass planet and
a ∼3000 d-period gas giant with HARPS (Lo Curto et al.
2010). In April 2010, we used Spitzer to search for the tran-
sit of the close-in planet HD 125612 c. Our initial light curve
showed a structure consistent with a transit, so we trig-
gered new observations of the transit window with Spitzer
in September 2010 that did not confirm the transit signal.
Reanalyzing the April data, we noticed that the transit sig-
nal disappeared when we included terms in the PSF widths,
revealing that the signal originated from the Spitzer PSF
breathing effect (Lanotte et al. 2014). Our global analysis of
the two Spitzer light curves allowed us to discard a transit-
ing configuration for the planet for any possible planetary
composition (Fig. 16), the resulting posterior probability
for a full transit being of only 0.24%, for a prior probability
of 9.7% (Table 1). We note that the standard deviations of
our Spitzer light curves binned per 30 min intervals are the
largest for HD 125612 (92 and 102 ppm, see Fig. 16). This
is also the case for the RV jitter measured from the HARPS
RVs (3.2 m s−1 ), suggesting that HD 125612 is significantly
more active star than our other targets.
M. Gillon et al.: Results of the Warm Spitzer transit search
5.16. HD 134060
HD 134060 is a G0-type dwarf with a V -magnitude of 6.3 at
24pc from Earth, for which M11 and Udry et al. (in prep.)
reported the discovery by HARPS of a Neptune (minimal
mass = 11 M⊕) on a 3.3d orbit, HD 134060 b, in addition
to a giant planet on a much wider orbit. Interestingly, the
HARPS RVs showed that HD 134060 b has a significant ec-
centricity of 0.40 ± 0.04, despite its very short orbit, rem-
iniscent of the still poorly understood eccentricity of the
hot Neptune prototype GJ 436 b (Lanotte et al. 2014). Our
analysis of the updated HARPS dataset confirmed this sig-
nificant eccentricity. The orientation of the elliptic orbit of
the planet made its occultation by its host star much more
likely than its transit (23 vs 8%, see Table A.4), and its tim-
ing much better constrained. We thus decided to use a few
hours of Spitzer first to try to detect the occultation of the
planet. We choose to perform the measurement at 3.6µm,
based on the clear detection of the occultation of GJ 436 b
at 3.6 µm and its non-detection at 4.5 µm (Stevenson et
al. 2010, Lanotte et al. 2014). Assuming black-body spec-
tra for the star and the planet, we estimated the expected
occultation depth to range from a few tens to more than
200ppm, depending on the unknown planet's size, albedo,
and heat-distribution efficiency. The resulting light curve
(Fig. 17, bottom panel) did not reveal a single eclipse, but
its precision was not high enough to discard an occultation
of the planet for a large range of the plausible planetary pa-
rameters cited above. We then attempted a transit search,
this time at 4.5 µm. The resulting light curve was flat (Fig.
17, top panel) and enabled us to fully discard a full transit
of the planet (Table 1).
5.17. HD 181433
Three planets were detected by HARPS around this V =
8.4 K-dwarf (Bouchy et al. 2008). Our analysis of the up-
dated HARPS dataset confirmed the existence of these
three planets, and resulted in a prior transit probabil-
ity of 5% for the inner most of them, the super-Earth
HD 181433 b (Mp sin i ∼ 7.5M⊕, P =9.375d). Our Spitzer
observations of one of its transit windows led to a transit
light curve that did not reveal any transit signature (Fig.
18), while being precise enough to discard a transit for any
possible composition of the planet. Our global MCMC anal-
ysis of the RVs and Spitzer photometry led to a posterior
full transit probability of 0.52% (Table 1), which is proba-
bly large enough to justify a future exploration of the first
part of the transit window that was left unexplored by our
Spitzer observations (see Fig. 18).
5.18. HD 215497
Two planets were detected by HARPS around this nearby
(∼44 pc) K-dwarf (Lo Curto et al. 2010), one of them be-
ing a giant planet on a significantly eccentric 568d-orbit,
and the other being a close-in (P =3.93d) super-Earth
(Mp sin i = 6M⊕). Our analysis of the updated HARPS
dataset, containing only three additional RVs, confirmed
the existence of both planets. Based on its interestingly
high transit probability of ∼ 12% the inner super-Earth
HD 215497 b, we targeted it in our Spitzer transit search.
The resulting light curve did not reveal any transit signa-
ture and was precise enough to discard a transit for most
Fig. 17. T op: same as Fig. 1 for HD 134060 b. Bottom: same
for the planet's occultation. The red solid line and blue
dashed line show, respectively, models for a central occul-
tation of a 1.85 R⊕ (Earth-like composition) and 5 R⊕ (H-
dominated composition) planet, assuming a null albedo, an
inefficient heat distribution to the night side, and negligible
tidal effects for the planet, and assuming black-body emis-
sions for both the planet and its star. The time range of
the x-axes of both panels correspond to the same duration
so as to outline the fact that the transit timing was much
less constrained by the RVs than the occultation timing.
Fig. 18. Same as Fig. 1 for HD 181433 b.
11
M. Gillon et al.: Results of the Warm Spitzer transit search
Fig. 19. Same as Fig. 1 for HD 215497 b.
plausible compositions. Yet, a high impact parameter tran-
sit of an iron-dominated planet is not discarded by the data
(see Fig. 19), resulting in a small but significant posterior
full transit probability of ∼ 0.31% (Table 1).
5.19. HD 219828
Discovered by Melo et al. in 2007, HD 219828 b is a hot
Neptune (Mp sin i = 20M⊕) with an interestingly high ge-
ometric transit probability of ∼14%, thanks to its close-in
orbit (a=0.05 au, P =3.83d), combined with the relatively
large size of its evolved G0-type host star (R∗ = 1.6R⊙). On
the other hand, this large stellar size translated in expected
transit depths as small as 100 ppm (pure iron composition),
so we monitored two transit windows to reach a photomet-
ric precision that was high enough to firmly constrain the
(non-)transiting nature of the planet. The resulting light
curves, both obtained at 4.5 µm, are shown in Fig. 20. They
do not reveal any transit signature. Our global analysis of
the RVs + photometry led to a complete rejection of a full
transit configuration (Table 1-. Our analysis of the much ex-
tended HARPS dataset compared to the HD 219828 b dis-
covery paper (91 vs 22 RVs) confirmed the existence of a
second, more massive planet (Mp sin i ∼ 15MJup) planet
on a very eccentric (e = 0.81) long-period (P =13.1 years)
orbit, as recently announced by Santos et al. (2016).
6. Discussion and conclusion
The constraint brought by the Spitzer photometry on the
transiting nature of a given planet can be directly esti-
mated in Table 1 by comparing the derived prior and pos-
terior transit probabilities. For 16 out of the 19 RV planets
targeted here, our Spitzer observations explored the tran-
sit window with a coverage and precision high enough to
make a transiting configuration very unlikely, the poste-
rior transit probabilities for these 16 targets all being less
than 0.55% (Table 1). For HD 45184 b, this posterior transit
probability is still of 1.3%, because our Spitzer observations
did not explore the second part of its transit window (see
Sect. 5.9 and Fig. 10). The transiting nature of HD 13808 b
and HD 115617 b is left unexplored by our observations.
For the radius of each planet, our MCMC analysis as-
sumed a uniform prior PDF ranging from a pure-iron com-
12
Fig. 20. Same as Fig. 1 for HD 219828 b.
position radius to 11 R⊕. The fact that a transiting config-
uration was disfavored by our MCMC analysis for all tar-
geted planets does not preclude the possibility that one or
several transits were in the data, but were just too shallow
to be noticed by the Markov Chains. To estimate the ac-
tual detection threshold of our observations, we performed
the following procedure for each of our targets. We created
50 fake transit light curves based on the multiplication of
the actual light curve by a transit model of the targeted
RV planets, each transit assuming a circular orbit for the
planet, a mid-transit time drawn from the prior transit tim-
ing PDF, which was derived from the RV analysis, and an
impact parameter drawn from a uniform PDF that ranged
from 0 to 0.9. For each transit model, the depth was au-
tomatically tuned to have a difference in BIC of +9.2 be-
tween models neglecting and taking into account the tran-
sit. This difference in BIC corresponds to a Bayes factor of
e9.2/2 = 100, indicating a decisive selection of the transit
model (Jeffreys 1961), i.e. a firm detection (at better than
3-3.5 sigma) of the transit. We then averaged the transit
depths (and planet's radii) derived for the 50 light curves
and adopted the resulting value as the detection threshold
for the considered dataset. These detection thresholds are
given in Table 1, expressed as transit depth (in ppm) and
planetary radius. For each planet, they are compared to the
planet's radius assuming pure-iron and Earth-like compo-
sitions. The detection threshold radius is smaller than the
ones assuming pure-iron and Earth-like compositions for,
M. Gillon et al.: Results of the Warm Spitzer transit search
respectively, 5 and 14 of the 17 planets for which the tran-
siting nature was constrained by our observations. For 12
and 3 planets, we cannot thus fully reject the absence of a
transit in our data, provided very metal-rich and Earth-like
compositions, respectively. However, no transiting configu-
ration is even midly favored for any planet by our MCMC
analysis (from the comparison of the last columns of Table
1), so, considering the excellent photometric precision of our
Spitzer data, the hypothesis of a missed transit is clearly
unlikely. Using the posterior transit probabilities shown in
a posterior probability that none of the probed 19 planets
transits of 83%, vs 22% for the corresponding prior proba-
Table 1, the formula 1 −Qi=1:19 Pi(tr, D) indeed results in
bility (1 −Qi=1:19 Pi(tr)).
Our multi-Cycle Spitzer transit search explored the
transiting nature of 25 RV planets. It detected one or
several transits for the planets HD 75732 e (aka 55 Cnc e)
(Demory et al. 2011, Gillon et al. 2012) and HD 219234 b
(Motalebi et al 2015), confirmed the transiting nature of
HD 97658 b (Van Grootel et al. 2015), discarded or disfa-
vored the transiting nature of 20 planets (including one
presented in S´egransan et al., in prep.), and left the one
of two planets unconstrained. By discovering the transits
of two planets of a few Earth-masses that are suitable for
detailed atmospheric characterization, it brought a signifi-
cant contribution to the study of super-Earths. Statistically
speaking, its final result is normal: considering all the plan-
ets listed in Table 1 except HD 97658 b, which we decided to
observe only because we knew that it was probably transit-
ing (Dragomir et al. 2013), the sum of the geometric transit
probabilities amounts to 196%, i.e. the project was expected
to observe the transits of ∼ 2 planets.
The photometric performances demonstrated by Spitzer
in this program are illustrated in Fig. 21. This figure com-
pares as a function of the targets' K-magnitude the stan-
dard deviations measured in the detrended light curves for a
sampling of 30 min to the corresponding formal errors com-
puted following the instructions of the Spitzer Observation
Manual (SOM6). At 4.5µm, the measured standard devia-
tions are well modeled by the linear relationship σ30min =
32.5 + 11.97 × (Kmag − 5) ppm, while the mean quadratic
difference between the measured standard deviations and
the formal errors is 35 ± 5 ppm. This quadratic difference
is 64 ± 9 ppm at 3.6 µm, and 45 ± 4 ppm when neglect-
ing HD125612, which seems to be a more active star than
our other targets (Sec. 5.15). These quadratic differences
can be attributed to the low-frequency noise of instrumen-
tal and astrophysical origins that cannot be represented by
our instrumental model. Figure 21 shows that this red noise
dominates the photometric precision of Spitzer, especially
for the brighter targets. Its values are low enough -- a few
dozens of ppm -- to qualify the photometric precision of
Spitzer as excellent, and to make it an optimal facility for
the search for very low-amplitude transits on bright nearby
stars.
The Spitzer mission should come to an end in early
2019. Fortunately, the CHEOPS space mission (Broeg et al.
2015) will arrive just in time to take over the search for the
transits of super-Earths discovered by RVs around nearby
stars. CHEOPS will not benefit from a targets' visibility
as favorable as Spitzer, but its full dedication to transit
6 http://ssc.spitzer.caltech.edu/warmmission/propkit/som/
Fig. 21. Standard deviations of the detrended Spitzer pho-
tometry binned per 30 min (open circles) and the cor-
responding formal errors (triangles) as a function of the
K-magnitude of the targets. Blue = 3.6 µm, red = 4.5
µm. The dashed red line shows the best-fit linear rela-
tionship between the standard deviations measured at 4.5
µm and the K-magnitudes, its equation being σ30min =
32.5 + 11.97 × (Kmag − 5) ppm.
observations will more than compensate for its geocentric
orbit.
Acknowledgements. This work is based in part on observations made
with the Spitzer Space Telescope, which is operated by the Jet
Propulsion Laboratory, California Institute of Technology under a
contract with NASA. Support for this work was provided by NASA.
M. Gillon is Research Associate at the Belgian Scientific Research
Fund (F.R.S-FNRS). This publication makes use of data products
from the Two Micron All Sky Survey, which is a joint project of
the University of Massachusetts and the Infrared Processing and
Analysis Center/California Institute of Technology, funded by the
National Aeronautics and Space Administration and the National
Science Foundation.
References
Arriagada, P., Anglada-Escud´e, G., Butler, R. P., et al. 2013, ApJ,
771, 42
Baranne, A., Queloz, D., Mayor, M. et al. 1996, A&AS, 119, 373
Bonfils, X., Gillon, M., Forveille, T., et al. 2011, A&A, 528, A111
Bouchy F., Mayor M., Lovis C., et al., 2009, A&A, 496, 527
Broeg, C., Fortier, A., Ehrenreich, D., et al. 2013, Hot Planets and
Cool Stars, Garching, Germany, Edited by Roberto Saglia; EPJ
Web of Conferences, Volume 47
Butler, R. P., Vogt, S. S., Marcy, G. W., et al. 2004, ApJ, 617, 580
Casagrande, L., Schoenrich, R., Asplund, M., et al. 2011, A&A, 530,
138
Charbonneau, D., Berta, Z. K., Irwin, J., et al. 2009, Nature, 462, 891
Claret, A. 2000, A&A, 363, 1081
Claret, A., & Bloemen, S. 2011, A&A, 529, A75
Demory, B.-O., Gillon, M., Deming, D., et al. 2011, A&A, 533, A114
Demory, B.-O., Gillon, M., Seager, S., et al. 2012, ApJL, 751, 28
D´ıaz, R. F., S´egransan, D., Udry, S., et al. 2016, A&A, 585, A134
Dragomir, D., Matthews, J. M., Eastman, J. D., et al. 2013, ApJL,
772, 2
13
M. Gillon et al.: Results of the Warm Spitzer transit search
Planet
Ms
d
[M⊙]
[pc]
K Mp sin i Detection
threshold
ppm/[R⊕]
[M⊕]
Pure-Fe Earth-like
radius
[R⊕]
radius
[R⊕]
HD 40307 b1
GJ 3634 b2
HD 75732 e3,4
HD 97658 b5
HD 219134 b6
BD-061339 b7
HD1461 b7
HD 10180 c7
HD 13808 b7
HD 20003 b7
HD 20781 b7
HD 31527 b7
HD 39194 b7
HD 45184 b7
HD 47186 b7
HD 51608 b7
HD 93385 d7
HD 96700 b7
HD 115617 b7
HD 125612 c7
HD 134060 b7
HD 181433 b7
HD 215497 b7
HD 219828 b7
0.78
0.45
0.91
0.77
0.78
0.63
1.04
1.06
0.77
0.91
0.83
0.96
0.72
1.00
1.03
0.86
1.04
0.96
0.94
1.09
1.07
0.86
0.87
1.18
12.8
19.8
12.3
21.1
6.5
20.3
23.2
39.0
28.6
43.8
35.4
38.6
25.9
21.9
39.6
34.8
42.2
25.7
8.6
54.2
24.2
26.8
43.6
72.3
4.79
7.47
4.02
5.73
5.57
6.31
4.90
5.87
6.78
6.65
6.55
6.05
6.09
4.87
6.01
6.33
6.07
5.00
2.96
6.84
4.84
6.09
6.78
6.53
4.3
7.0
7.8
7.6
4.3
6.9
6.7
13.1
11.8
11.8
2.1
10.7
4.1
12.1
23.2
13.1
4.0
9.1
6.2
19.3
10.0
7.5
6.1
20.2
150/0.91
500/1.05
NAa
NAa
NAa
235/1.15
170/1.56
180/1.73
NAa
280/1.79
220/1.39
190/1.64
210/1.22
150/1.39
175/1.59
195/1.39
200/1.78
155/1.58
NAa
290/1.92
165/1.61
190/1.26
250/1.66
155/2.18
1.13
1.31
1.34
1.33
1.15
1.30
1.29
1.53
1.50
1.49
0.93
1.46
1.12
1.50
1.77
1.54
1.11
1.40
1.26
1.69
1.43
1.33
1.26
1.71
1.44
1.65
1.70
1.69
1.44
1.64
1.63
1.95
1.90
1.90
1.17
1.85
1.42
1.91
2.26
1.95
1.41
1.77
1.60
2.16
1.82
1.68
1.59
2.19
Prior
P (tr)
Posterior
P (tr, D)
[%]
6.6
7.0
28.9
4.3
9.5
7.7
8.1
8.4
3.5
3.4
7.1
4.4
6.4
7.7
10.5
4.0
7.9
6.8
9.4
9.7
8.4
4.9
11.8
14.2
[%]
0.19
0.50
100
100
100
0.53
0.14
0.14
3.5
0.54
0.15
0.46
0.46
1.29
0.00
0.11
0.23
0.09
9.4
0.24
0.00
0.52
0.31
0.00
Table 1: 24 of the 25 planets targeted by our Spitzer multi-cycle transit search. The first group of 5 planets are the targets
for which results were presented previously. The second group of 19 planets are the targets of this work. For each planet,
column 6 provides the mean transit detection threshold (in ppm and [R⊕]), as inferred from injection of transit models
in the data (see Sec. 6 for details). Column 7 and 8 give the theoretical radii corresponding, respectively, to a pure-iron
and Earth-like compositions. Columns 9 and 10 present the a priori (geometric) and a posteriori (after observation)
transit probabilities for the planet. aNot Applicable: for HD 13808 b and HD 115617 b (61 Vir b), our observations did
not constrain the (non-)transiting configuration, while for HD 75732 e (55 Cnc e), HD 97658 b, and HD 219134 b, a transit
was firmly detected in the data. References: 1: Gillon et al. 2010, 2: Bonfils et al. 2011, 3: Demory et al. 2011, 4: Gillon
et al. 2012a, 5: Van Grootel et al. 2014, 6: Motalebi et al. 2015, 7: this work.
Dumusque, X., Pepe, F., Lovis, C., & Latham, D. W., 2015, ApJ, 808,
171
Hog, E., Fabricius, C., Makarov, V. V., et al. 2000, A&A, 355, 27
Houk, N. & Cowley, A. P. 1975, Michigan Spectral Survey, Ann Arbor,
Ehrenreich, D., Bourrier, V., Wheatley, P. J., et al. 2015, Nature, 522,
Dep. Astron., Univ. Michigan, 1
459
Howard, A. W., Marcy, G. W., Bryson, S. T., et al. 2012, ApJS, 201,
Fazio, G. G., Hora, J. L., Allen, L. E., et al., 2004, ApJS, 154, 10
Fischer, D., Marcy, G., Butler, P., et al. 2007, ApJ, 669, 1336
Flower, P. J. 1996, ApJ, 469, 355
Gelman, A., Rubin, D. 1992, Statist. Sci., 7, 457
Gillon M., Pont F., Demory B.-O., et al, 2007, A&A, 472, L13
Gillon, M., Deming, D., Demory, B.-O., et al. 2010, A&A, 518, A25
Gillon, M., Demory, B.-O., Benneke, B., et al. 2012a, A&A, 539, A28
Gillon, M., Triaud, A. H. M. J., Fortney, J. J., et al. 2012b, A&A,
15
Ingalls, J. G., Krick J. E., Carey, S. J., et al. 2012, Proc. SPIE 8442
Jeffreys, H. 1961, The Theory of Probability, 3rd ed. Oxford University
Press.
Kammer, J. A., Knutson H. A., Howard, A. W., et al. 2014, ApJ, 781,
103
Kharchenko, N. V. 2001, Kinematika i Fizika Nebesnykh Tel, 17, 409
Knutson, H. A., Charbonneau, D., Allen, L. A. et al., 2008, ApJ, 673,
542, A4
526
Gillon, M., Demory, B.-O., Madhusudhan, N., et al. 2014, A&A, 563,
Knutson, H. A., Benneke, B., Deming, D. & Homeier, D. 2014a,
A21
Nature, 505, 66
Glebocki, R. & Gnacinski, P. 2005, Catalogue of Stellar Rotational
Knutson, H. A., Dragomir, D., Kreidberg, L., et al. 2014b, ApJ, 794,
Velocities, ESA, SP-560, 571
155
Gray, R. O., Corbally, C. J., Garrison, R. F., et al. 2003, AJ, 126,
2048
Gray, R. O., Corbally, C. J., Garrison, R. F., et al. 2006, AJ, 132, 161
Kreidberg, L., Bean, J., Desert, J.-M., et al., 2014, Nature, 505, 69
Lanotte, A. A., Gillon, M., Demory, B.-O., et al. 2014, A&A, 572, 73
Lo Curto G., Mayor, M., Benz, W., et al. 2010, A&A, 512, A48
14
M. Gillon et al.: Results of the Warm Spitzer transit search
Lo Curto, G., Mayor, M., Benz, W., et al. 2013, A&A, 551, 59
Lanotte, A., Gillon, M., Demory, B.-O., et al. 2014, A&A, 572, 73
Lovis, C., S´egransan, D., Mayor, M., et al. 2011, A&A, 528, A112
Mandel, J., & Agol, E. 2002, ApJ, 580, 171
Mayor, M., Pepe, F., Queloz, D., et al. 2003, The Messenger, 114, 20
Mayor, M., Marnier, M., Lovis, C., et al. 2011, A&A (submitted),
arXiv1109.2497
Melo, C., Santos, N. C., Gieren, W., et al. 2007, A&A, 467, 721
Meschiari, S., Wolf, A. S., Rivera, E. et al., 2009, PASP, 121, 1016
Motalebi, F., Udry, S., Gillon, M., et al. 2015, A&A, 584, A72
Noyes, R. W., Hartmann, L. W., Baliunas, S. L., et al. 1984, ApJ,
279, 763
Press, W. H., Teukolsky, S. A., Vetterling, W. T., Flannery, B. P. 1992,
Numerical Recipes in Fortan 77: The Art of Scientific Computing,
Cambridge Universty Press
Queloz, D., Henry, G. W., Sivan, J. P., et al. 2001, A&A, 379, 279
Rafikov R. R., 2006, ApJ, 648, 666
Rivera, E. J, Butler, R. P., Vogt, S. S., et al. 2010, ApJ, 708, 1492
Rowe, J. F., Bryson, S. T., Marcy, G. W., et al. 2014, ApJ, 784, 45
Santos, N. C., Bouchy, F., Mayor, M., et al. 2004, A&A, 426, L19
Santos, N. C., Sousa, S. G., Mortier, A., et al. 2013, A&A, 556, A150
Santos, N. C., Santerne, A., Faria, J. P., et al. 2016, A&A, 592, A13
Schwarz, G. E. 1978, Annals of Statistics, 6, 461
Seager, S., Kuchner, M., Hier-Majumder, C. A., Militzer, B. 2007,
ApJ, 669, 1279
Seager, S. 2010, Exoplanet Atmospheres, Princeton University Press
Skrutskie, M. F., Cutri, R. M., Stiening, R., et al. 2006, AJ, 131, 1163
Stevens, D. J. & Gaudi, B. S. 2013, PASP, 125, 933
Stevenson, K. B., Harrington, J., Nymeyer, S., et al. 2010, Nature,
464, 1161
Stevenson, K. B., Harrington, J., Fortney, J., et al. 2012, ApJ, 754,
136
Stetson, P. B. 1987, PASP, 99, 111
Valencia D., Guillot, T., Parmentier, V., Freedman, R. S. 2013, ApJ,
775, 10
Van Grootel, V., Gillon, M., Valencia, D., et al. 2014, ApJ, 786, 2
Van Leeuwen, F. 2007, A&A, 474, 653
Vogt, S. S., Wittenmeyer, R. A., Butler, R. P., et al. 2010, ApJ, 708,
1366
Wisniewski, J. P., Ge, J., Crepp, J. R., et al. 2012, AJ, 143, 107
Winn, J. N., Matthews, J. M., Dawson, R. I., et al. 2011, ApJ, 733,
L18
Wright, J. T. 2005, PASP, 117, 657
Appendix A: Tables
The following tables describe the targets of this work
(Tables A.1 to A.4) and the Spitzer observations (Tables
A.5 to A.11).
15
M. Gillon et al.: Results of the Warm Spitzer transit search
BD-061339
20.3 ± 0.7(1)
9.67 ± 0.04(2)
6.31 ± 0.02(3)
K7V/M0V(4)
4040 ± 50(5)
−0.07 ± 0.10(5)
0.63 ± 0.03(5)
−1.07 ± 0.03(6)
0.69 ± 0.02
4.56 ± 0.04
HD 1461
HD 10180
HD 13808
HD 20003
23.2 ± 0.3(1)
6.46 ± 0.01(2)
4.90 ± 0.02(3)
G3V(7)
5765 ± 50(8)
0.19 ± 0.01(8)
1.04 ± 0.07(8)
39.0 ± 0.6(1)
7.32 ± 0.01(12)
5.87 ± 0.02(3)
G1V(13)
5910 ± 50(14)
0.08 ± 0.01(14)
1.06 ± 0.05(14)
28.6 ± 0.5(1)
8.38 ± 0.01(2)
6.25 ± 0.02(3)
K2V(13)
5035 ± 50(8)
−0.21 ± 0.02(8)
0.77 ± 0.06(8)
43.8 ± 1.2(1)
8.37 ± 0.01(2)
6.65 ± 0.02(3)
G8V(16)
5495 ± 50(8)
0.04 ± 0.02(8)
0.91 ± 0.07(8)
−0.082 ± 0.009(6)
−0.058 ± 0.009(6)
−0.290 ± 0.021(6)
−0.140 ± 0.013(6)
1.10 ± 0.02
4.37 ± 0.04
1.18 ± 0.02
4.32 ± 0.03
0.81 ± 0.02
4.51 ± 0.05
0.98 ± 0.02
4.41 ± 0.04
HARPS: 102(4)+10
HARPS: 167(9)+82(10)+5
HARPS: 190(14)+63
HARPS: 133(9)+89
HARPS: 104(9)+77
Star
d [parsec]
V [mag]
K [mag]
Spectral type
Tef f [K]a
[Fe/H] [dex]
M∗ [M⊙]
Bolometric Correction
R∗ [R⊙]b
log g [cgs]
RV
Data
Model
PFS: 15(5)
2 Keplerians
Keck: 144(11)
3 Keplerians
+ linear trend
+ quadratic trend
Jitter noise [m s−1 ]
HARPS: 2.7
PFS: 2.7
+CCF function
HARPS: 1.5
Keck: 2.4 & 1.6e
6 Keplerians
2 Keplerians
4 Keplerians
+ quadratic trend
+CCF function
1.5
1.7
1.4
Planet
BD-061339 b(4)
HD 1461 b(11)
HD 10180 c(14)
HD 13808 b(9,15)
HD 20003 b(9,17)
Mp sin i [M⊕]
Min. Rp [R⊕]c
Min. (Rp/R∗)2 [ppm]c
Teq [K]d
T0-2450000 [BJDT DB ]
6.93 ± 0.96
1.30 ± 0.05
298 ± 29
796 ± 17
6627.48+0.18
−0.16
6.73 ± 0.47
1.29 ± 0.03
115 ± 6
1154 ± 20
6549.30 ± 0.12
13.11 ± 0.62
1.53 ± 0.02
142 ± 6
1223 ± 18
5212.837+0.059
−0.074
11.83 ± 0.88
1.50 ± 0.03
286 ± 18
674 ± 14
11.79 ± 0.61
1.49 ± 0.02
191 ± 9
836 ± 12
6537.49 ± 0.26
6538.34 ± 0.36
P [d]
Wb=0 [min]
K [m s−1 ]
a [AU]
e
ω [deg]
Prior Ptransit [%]
Prior Poccultation [%]
3.87310 ± 0.00037
5.77198 ± 0.00030
5.75931 ± 0.00021
14.1853 ± 0.0019
11.8489 ± 0.0015
139 ± 15
3.90 ± 0.52
218 ± 11
2.34 ± 0.13
238 ± 10
4.51 ± 0.15
246 ± 13
3.73 ± 0.20
388 ± 23
3.84 ± 0.20
0.04136 ± 0.00066
0.0638 ± 0.0015
0.0641 ± 0.0010
0.1050 ± 0.0028
0.09817 ± 0.00072
0.11+0.11
−0.08
192+69
−93
7.72+0.77
−0.71
7.9+1.1
−0.7
0.037+0.041
−0.026
134+110
−120
8.06 ± 0.38
8.00 ± 0.38
0.045+0.037
−0.030
320+51
−41
8.38+0.30
−0.33
8.78+0.38
−0.32
0.042+0.043
−0.029
272+85
−75
3.52 ± 0.18
3.68+0.21
−0.18
0.377 ± 0.047
267.1+7.6
−8.3
3.43 ± 0.15
7.50+0.62
−0.56
Table A.1: Targets 1-5 of our Warm Spitzer program. For each target, the table presents the assumed stellar parameters,
the RV data and details of their analysis, and the parameters that we derived for the planet searched for transit from our
RV analysis. aFor the sake of realism, a minimal error of 50K was assumed here. bFrom luminosity and Tef f . cAssuming
Mp sin i = Mp. These minimum values correspond to a pure iron planet (Seager et al. 2007). dAssuming a null albedo
and a heat distribution factor f ′ = 1/4 (Seager 2010). eFor the first and second Keck datasets presented in Rivera et
al. (2010), respectively. References: (1)Van Leeuwen (2007), (2)Kharchenko (2001), (3)Skrutskie et al. (2006), (4)Lo Curto
et al. (2013), (5)Arriagada et al. (2013), (6)Flower (1996), (7)Gray et al. (2003), (8)Santos et al. (2013), (9)Mayor et al.
(2011), (10)D´ıaz et al. (2015), (11)Rivera et al. (2010), (12)Hog et al. (2000), (13)Gray et al. (2006), (14)Lovis et al. (2011),
(15)Queloz et al. in prep., (16)Houk & Cowley (1975), (17)Udry et al. in prep.
16
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
d [parsec]
V [mag]
K [mag]
Spectral type
Tef f [K]a
[Fe/H] [dex]
M∗ [M⊙]
HD 20781
HD 31527
HD 39194
HD 45184
35.4 ± 1.3(1)
8.44 ± 0.01(2)
6.55 ± 0.02(3)
K0V(4)
5255 ± 50(5)
−0.11 ± 0.02(5)
0.83 ± 0.06(5)
38.6 ± 0.9(1)
7.48 ± 0.01(2)
6.05 ± 0.02(3)
G0V(9)
5900 ± 50(5)
-0.17 ± 0.01(5)
0.96 ± 0.07(5)
21.9 ± 0.2(1)
8.08 ± 0.01(2)
6.09 ± 0.02(3)
K0V(4)
5205 ± 50(5)
−0.61 ± 0.02(5)
0.72 ± 0.05(2)
6.37 ± 0.01(4)
4.87 ± 0.02(3)
G2V(4)
5870 ± 50(5)
0.04 ± 0.01(5)
1.00 ± 0.07(5)
Bolometric Correction
−0.208 ± 0.016(6)
−0.059 ± 0.009(6)
−0.225 ± 0.017(6)
−0.058 ± 0.008(6)
R∗ [R⊙]b
log g [cgs]
RV
Data
Model
0.86 ± 0.02
4.49 ± 0.04
1.09 ± 0.02
4.35 ± 0.04
0.77 ± 0.02
4.52 ± 0.04
1.04 ± 0.02
4.40 ± 0.04
HARPS: 96(7)+117
HARPS: 167(7)+75
HARPS: 133(7)+128
HARPS: 82(7)+92
4 Keplerians
3 Keplerians
3 Keplerians
2 Keplerians
+ quadratic trend
+ quartic trend
+ CCF function
+ CCF function
Jitter noise [m s−1 ]
1.2
1.2
1.1
1.95
Planet
HD 20781 b(8)
HD 31527 b(7,8)
HD 39194 b(7,9)
HD 45184 b(7,8)
Mp sin i [M⊕]
Min. Rp [R⊕]c
Min. (Rp/R∗)2 [ppm]c
Teq [K]d
2.12 ± 0.35
0.93 ± 0.04
99 ± 11
993 ± 20
10.68 ± 0.71
1.46 ± 0.03
150 ± 8
839 ± 15
4.08 ± 0.32
1.12 ± 0.03
179 ± 13
935 ± 19
12.08 ± 0.86
1.50 ± 0.03
175 ± 10
1143 ± 20
T0-2450000 [BJDT DB ]
6613.92 ± 0.23
6792.65 ± 0.36
6285.93 ± 0.15
6317.67 ± 0.13
P [d]
Wb=0 [min]
K [m s−1 ]
a [AU]
e
ω [deg]
5.3144 ± 0.0011
16.5547 ± 0.0033
5.63675 ± 0.00044
5.88607 ± 0.00032
180 ± 6
0.88 ± 0.14
293 ± 19
2.78 ± 0.13
174 ± 10
1.83 ± 0.12
209 ± 13
4.33 ± 0.23
0.0560 ± 0.0014
0.1253 ± 0.0031
0.0555 ± 0.0013
0.0638 ± 0.0015
0 (fixed)
0.117 ± 0.052
-
Prior Ptransit [%]
7.14 ± 0.25
Prior Poccultation [%]
7.14 ± 0.25
42+23
−26
4.40 ± 0.29
3.81 ± 0.22
0.033+0.035
−0.023
224+97
−90
6.38 ± 0.36
6.57+0.42
−0.33
0.122+0.053
−0.057
178+30
−27
7.72+0.51
−0.44
7.67+0.48
−0.43
Table A.2: Same as Table A.1 for targets 6-9. References: (1)Van Leeuwen (2007), (2)Hog et al. (2000), (3)Skrutskie et
al. (2006), (4)Kharchenko (2001), (5)Santos et al. (2013), (6)Flower (1996), (7)Mayor et al. (2011), (8)Udry et al. in prep.,
(9)Queloz et al. in prep.
17
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
d [parsec]
V [mag]
K [mag]
Spectral type
Tef f [K]a
[Fe/H] [dex]
M∗ [M⊙]
HD 47186
HD 51608
HD 93385
HD 96700
HD 115617
39.6 ± 1.0(1)
7.63 ± 0.01(2)
6.01 ± 0.03(3)
G5V(4)
5675 ± 50(5)
0.23 ± 0.02(5)
1.03 ± 0.07(5)
34.8 ± 0.7(1)
8.17 ± 0.01(2)
6.33 ± 0.02(3)
G7V(4)
5360 ± 50(5)
-0.07 ± 0.01(5)
0.86 ± 0.06(5)
42.2 ± 1.3(1)
7.49 ± 0.01(2)
6.07 ± 0.02(3)
G2/G3V(4)
5975 ± 50(5)
0.02 ± 0.01(5)
1.04 ± 0.07(5)
25.7 ± 0.4(1)
6.50 ± 0.01(2)
5.00 ± 0.02(3)
G1/G2V(4)
5845 ± 50(5)
−0.18 ± 0.01(5)
0.96 ± 0.07(5)
8.56 ± 0.02(1)
4.73 ± 0.01(4)
2.96 ± 0.24(3)
G5V(4)
5575 ± 50(5)
0.01 ± 0.05(5)
0.94 ± 0.08(5)
Bolometric Correction
−0.100 ± 0.010(6)
−0.176 ± 0.016(6)
−0.049 ± 0.007(6)
−0.069 ± 0.008(6)
−0.120 ± 0.012(6)
1.10 ± 0.02
4.37 ± 0.04
0.91 ± 0.02
4.45 ± 0.04
1.15 ± 0.02
4.33 ± 0.04
1.16 ± 0.02
4.29 ± 0.04
0.98 ± 0.02
4.43 ± 0.05
HARPS: 66(7)+67
HARPS: 118(8)+92
HARPS: 127(8)+106
HARPS: 146(8)+98
Model
2 Keplerians
2 Keplerians
3 Keplerians
2 Keplerians
+ quadratic trend
+ CCF function
+ quadratic trend
+ CCF function
Jitter noise [m s−1 ]
0.9
1.2
1.3
1.6
AAT: 126(10)
+ Keck: 80(10)
3 Keplerians
Keck: 2.3
AAT: 2.2
R∗ [R⊙]a
log g [cgs]
RV
Data
Wb=0 [min]
K [m s−1 ]
a [AU]
e
ω [deg]
Planet
HD 47186 b(7)
HD 51608 b(8,9)
HD 93385 d(10)
HD 96700 b(7,10)
HD 115617 b(11)
Mp sin i [M⊕]
Min. Rp [R⊕]c
Min. (Rp/R∗)2 [ppm]c
Teq [K]d
23.2 ± 1.1
1.77 ± 0.02
217 ± 9
1277 ± 22
T0-2450000 [BJDT DB ]
5179.972 ± 0.021
13.12 ± 0.77
1.54 ± 0.02
239 ± 13
749 ± 14
6379.95+0.18
−0.16
3.97 ± 0.48
1.11 ± 0.03
79 ± 7
1129 ± 19
9.05 ± 0.64
1.40 ± 0.02
121 ± 6
1087 ± 19
5.86 ± 0.66
1.24 ± 0.04
135 ± 9
1190 ± 24
6364.17 ± 0.27
6521.26 ± 0.16
5280.71 ± 0.17
P [d]
4.084575 ± 0.000043
14.0702 ± 0.0015
7.3422 ± 0.0014
8.12541 ± 0.00068
4.21504 ± 0.00061
190 ± 6
9.1 ± 1.3
255 ± 11
3.85 ± 0.14
227+20
−24
4.51 ± 0.15
273+16
−13
174 ± 9
2.97 ± 0.15
2.44 ± 0.24
0.0505 ± 0.0012
0.1084 ± 0.0025
0.0749 ± 0.0017
0.0780 ± 0.0019
0.0500 ± 0.0014
0.039 ± 0.014
57 ± 23
Prior Ptransit [%]
10.45 ± 0.34
Prior Poccultation [%]
9.84 ± 0.31
0.033+0.033
−0.023
130+74
−66
3.97+0.19
−0.16
3.84 ± 0.17
0.13+0.10
−0.09
87 ± 60
7.9+1.1
−0.7
6.68+0.54
−0.61
0.049+0.049
−0.034
293 ± 74
6.76+0.33
−0.36
7.12+0.45
−0.34
0.078+0.083
−0.055
58+73
−160
9.45+0.99
−0.63
8.89+0.63
−0.72
Table A.3: Same as Table A.1 for targets 10-14. References: (1)Van Leeuwen (2007), (2)Hog et al. (2000), (3)Skrutskie et
al. (2006), (4)Kharchenko (2001),(5)Santos et al. (2013), (6)Flower (1996), (7)Bouchy et al. (2009), (8)Mayor et al. (2011),
(9)Udry et al. in prep., (10)Queloz et al. in prep., (11)Vogt et al. (2010)
18
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
HD 125612
HD 134060
HD 181433
HD 215497
HD 219828
d [parsec]
V [mag]
K [mag]
Spectral type
Tef f [K]a
[Fe/H] [dex]
M∗ [M⊙]
54.2 ± 3.0(1)
8.32 ± 0.02(2)
6.84 ± 0.03(3)
G3V(4)
5915 ± 50(5)
0.24 ± 0.01(5)
1.09 ± 0.07(5)
24.2 ± 0.3(1)
6.29 ± 0.01(2)
4.84 ± 0.03(3)
G3IV(4)
5965 ± 50(5)
0.14 ± 0.01(5)
1.07 ± 0.07(5)
26.8 ± 0.8(1)
8.38 ± 0.01(4)
6.09 ± 0.02(3)
K5V(4)
4880 ± 50(5)
0.36 ± 0.18(5)
0.86 ± 0.17(5)
43.6 ± 2.0(1)
8.95 ± 0.02(2)
6.78 ± 0.02(3)
K3V(4)
5000 ± 100(5)
0.25 ± 0.05(5)
0.87 ± 0.11(5)
72.3 ± 3.9(1)
8.01 ± 0.01(2)
6.53 ± 0.02(3)
G0IV(4)
5890 ± 50(5)
0.19 ± 0.03(5)
1.18 ± 0.08(5)
Bolometric Correction
−0.057 ± 0.008(6)
−0.050 ± 0.007(6)
−0.360 ± 0.024(6)
−0.304 ± 0.048(6)
−0.061 ± 0.008(6)
R∗ [R⊙]a
log g [cgs]
RV
Data
Model
Jitter noise [m s−1 ]
1.03 ± 0.02
4.43 ± 0.03
1.15 ± 0.02
4.35 ± 0.04
0.84 ± 0.02
4.51 ± 0.11
0.97 ± 0.04
4.40 ± 0.08
1.60 ± 0.03
4.10 ± 0.04
HARPS: 58(7)+10
HARPS: 100(9)+50
HARPS: 107(11)+78
HARPS: 105(7) + 3
HARPS: 22(21)+69
Keck: 19(8)
3 Keplerians
HARPS: 3.2
Keck: 4.7
2 Keplerians
3 Keplerians
2 Keplerians
2 Keplerians
+ CCF & logR'(HK)
+ CCF function
function
1.3
1.0
1.3
1.5
Planet
HD 125612 c(7)
HD 134060 b(9,10)
HD 181433 b(11)
HD 215497 b(7)
HD 219828 b(12)
Mp sin i [M⊕]
Min. Rp [R⊕]c
Min. (Rp/R∗)2 [ppm]c
Teq [K]d
19.3 ± 2.1
1.69 ± 0.04
226 ± 14
1269 ± 21
T0-2450000 [BJDT DB ]
5296.79 ± 0.14
P [d]
4.15514 ± 0.00044
9.97 ± 0.60
1.43 ± 0.03
130 ± 8
1469 ± 24
6416.96+0.10
−0.09
3.269555+0.000092
−0.000080
309 ± 16
4.69 ± 0.19
Wb=0 [min]
K [m s−1 ]
a [AU]
e
ω [deg]
Prior Ptransit [%]
Prior Poccultation [%]
172+12
−16
7.33 ± 0.73
0.0520 ± 0.0011
0.0441 ± 0.0010
0.093+0.090
−0.064
120+66
−71
9.7+1.1
−0.7
8.86+0.59
−0.69
0.480 ± 0.034
258.5 ± 5.2
8.39 ± 0.32
23.2+1.7
−1.5
7.5 ± 1.1
1.33 ± 0.06
210 ± 18
753+29
−23
6.11 ± 0.78
1.26 ± 0.04
141 ± 15
1101 ± 39
20.2 ± 1.2
1.71 ± 0.02
96 ± 5
1596 ± 27
6265.33 ± 0.17
5393.96 ± 0.14
5410.658 ± 0.050
9.37518 ± 0.00056
3.93394 ± 0.00065
3.834863 ± 0.000094
227 ± 19
2.72 ± 0.13
0.0822+0.0050
−0.0058
0.380 ± 0.041
198.2 ± 7.1
4.93+0.49
−0.39
6.25+0.64
−0.51
152 ± 19
2.81 ± 0.27
271 ± 13
7.43 ± 0.27
0.0465 ± 0.0020
0.0507 ± 0.0012
0.215 ± 0.096
0.063 ± 0.036
122 ± 30
11.8+1.8
−1.4
8.57+0.97
−0.87
228 ± 39
14.18 ± 0.64
15.32+0.76
−0.67
Table A.4: Same as Table A.1 for targets 15-19. References: (1)Van Leeuwen (2007), (2)Hog et al. (2000), (3)Skrutskie
et al. (2006), (4)Kharchenko (2001),(5)Santos et al. (2013), (6)Flower (1996), (7)Lo Curto et al. (2010), (8)Fischer et al.
(2007), (9)Mayor et al. (2011), (10)Udry et al. in prep., (11)Bouchy et al. (2009), (12)Melo et al. (2007).
19
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
Program ID
BD -061339
90072
Observation date
2013-11-30
HD 1461
(1): 80220(1)
(2): 90072
(1): 2011-08-31
(2): 2013-09-13
Channel [µm]
4.5
4.5
AOR(s)a
(S1): 48815616
(1): 48815360
(2): 48815104
(1): 42790656
(S2): 48816128
(2): 48815872
Spitzer pipeline version
S19.1.0
Exposure time [s]
NBCD
b
Duration [hr]
0.4
(1):1038
(2): 970
(1): 7.9
(2): 7.4
S19.1.0
0.1
(1): 5548
(2): 3600
(1): 12.9
(2): 8.4
Photometric aperture [pixels]
(1) & (2): 3.5
(1) & (2): 3.5
HD 10180
(1): 60027
(2) & (3): 90072
(1): 2010-01-16
(2) : 2013-08-10
(3): 2013-09-09
(1): 3.6
(2) & (3): 4.5
(1): 38139392
(S2): 48596224
(2): 48595968
(S3): 48595712
(3): 48595456
(1): S18.18.0
(2) & (3): S19.1.0
0.1
(1): 4400
(2) & (3): 3175
(1): 10.4
(2) & (3): 7.4
(1): 3
(2): 2.5
Baseline modelc
(1): p(wx + [xy]2) +BM
(1): p(w2
x + wy3 + [xy]3 + l2) +BM (1): p(t + wx + w2
(3): 3
y + [xy]2 + l) + BM
(2): p(wx + wy + [xy]2) + BM
(2): p(wx + [xy]2) + BM
(2): p(wx + [xy]2 + l2) + BM
Error correction factorsd
(1): βw = 0.94, βr=1.21
(1): βw = 0.97, βr=1.91
(2): βw = 0.91, βr=1.24
(2): βw = 0.94, βr=1.34
(3): p(wx + wy + [xy]2 + l) + BM
(1): βw = 0.95, βr=1.62
(2): βw = 0.97, βr=1.26
(3): βw = 0.99, βr=1.42
Table A.5: Details of our Warm Spitzer data and their analysis for targets 1-3. aAOR = astronomical observation fequest
= Spitzer observing sequence. (S) designates a short pre-run AOR performed to stabilize the pointing and instrument.
bBCD = Basic Calibrated Data = block of 64 Spitzer/IRAC subarray exposures. cFor the baseline model, p(ǫN ) denotes,
respectively, an N -order polynomial function of the logarithm of time (ǫ = l), of the PSF x- and y-positions (ǫ = [xy]),
and widths (ǫ = wx & wy). BM=BLISS mapping. dsee Sec. 4. References: (1)Kammer et al. (2014).
20
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
Spitzer pipeline version
Exposure time [s]
NBCD
b
HD 13808
90072
(1) 2013-08-15
(2) 2013-08-27
4.5
(S1): 48814848
(1): 48814592
(S2): 48818176
(2): 48817920
S19.1.0
0.4
(1) & (2): 970
Duration [hr]
(1) & (2): 7.4
HD 20003
90072
2013-09-01
4.5
(S): 48408576
(1): 48408320
(2): 48408064
(3): 48407808
S19.1.0
0.4
(1) & (2): 1498
(3): 580
(1) & (2): 11.5
(3): 4.4
HD 20781
90072
2013-11-16
4.5
(S): 48817664
(1): 48817408
(2): 488817152
(3): 48816896
(1): S19.1.0
0.4
(1) & (2): 1367
(3): 580
(1) & (2): 10.4
(3): 4.4
Photometric aperture [pixels]
Baseline modelc
(1): 2.75
(2): 3.5
(1) & (2) & (3): 3.5
(1) & (2) & (3): 2.75
(1): p(wx + wy + [xy]2) +BM (1): p(w2
(2): p(w2
x + w2
y + [xy]) + BM (2): p(wx + wy + [xy]) + BM (2): p(w2
x + w2
y + [xy]2) +BM (1): p(wx + wy + [xy]2) + BM
x + wy + [xy]2) + BM
Error correction factorsd
(1): βw = 0.93, βr=1.00
(1): βw = 0.93, βr=1.30
(1): βw = 0.97, βr =1.06
(3): p(wx + wy + [xy]) + BM
(3): p(wx + [xy]2) + BM
(2): βw = 0.96, βr=1.55
(2): βw = 0.95, βr=1.12
(2): βw = 0.91, βr =1.14
(3): βw = 0.98, βr=2.03
(3): βw = 0.97 βr=1.31
Table A.6: Same as Table A.5 for targets 4-6.
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
Spitzer pipeline version
Exposure time [s]
NBCD
b
Duration [hr]
Photometric aperture [pixels]
Baseline modelc
HD 31527
90072
2014-05-13
4.5
(S): 50091776
(1): 50091520
(2): 50091264
(3): 50091008
S19.1.0
0.4
(1) & (2) & (3): 1558
(1) & (2) & (3): 11.9
(1) & (2) & (3) : 3.25
x + [xy]2) +BM
(1): p(w2
(2): p(wx + wy + [xy]2) + BM
y + [xy]2) + BM
(3): p(wx + w3
HD 39194
90072
2012-12-24
4.5
(S): 46914816
(1): 46914560
(1): S19.1.0
0.1
(1): 4890
(1): 11.4
(1): 3.0
(1): p(wx + wy + [xy]2 + l) + BM
Error correction factorsd
(1): βw = 0.93, βr=1.19
(1): βw = 0.97, βr=1.15
(2): βw = 0.95, βr=1.20
(3): βw = 0.93, βr=1.38
Table A.7: Same as Table A.5 for targets 7-8.
21
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
Spitzer pipeline version
Exposure time [s]
NBCD
b
Duration [hr]
HD 45184
90072
2013-01-24
4.5
(S): 46917376
(1): 46917120
S19.1.0
0.1
(1): 4880
(1): 11.4
Photometric aperture [pixels]
(1): 3.25
HD 47186
60027
2009-12-14
3.6
38065664
S19.1.0
0.1
2115
5
2.75
HD 51608
90072
2013-03-27
4.5
(S): 48411392
(1): 48411136
(2): 48410880
(1): S19.1.0
0.4
(1): 1498
(2): 1300
(1): 11.4
(2): 9.9
(1) & (2): 2.25
Baseline modelc
(1): p(w2
x + wy + [xy]2 + l) +BM p(w2
x + [xy]2 + l) +BM (1): p(wx + wy + [xy]2) + BM
(2): p(wx + wy + [xy]2) + BM
Error correction factorsd
(1): βw = 0.94, βr=1.54
(1): βw = 1.00, βr=1.61
(1): βw = 0.95, βr=1.14
(2): βw = 0.94, βr=1.12
Table A.8: Same as Table A.5 for targets 9-11.
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
Spitzer pipeline version
Exposure time [s]
NBCD
b
Duration [hr]
HD 93385
90072
2013-03-12
4.5
(S): 48407552
(1): 48407296
(2): 48407040
(3): 48406784
S19.1.0
0.1
(1) & (2): 4245
(3): 2325
(1) & (2): 11.9
(3): 5.4
HD 96700
90072
2013-08-16
4.5
(S): 46916864
(1): 46916608
(2): 46916352
S19.1.0
0.1
(1): 4880
(2): 1240
(1): 11.4
(2): 2.9
HD 115617
60027
2010-03-25
3.6
(1): 39138816
(2): 39139072
S19.1.0
0.01
(1): 6600
(2): 6640
(1): 6.2
(2): 6.3
Photometric aperture [pixels]
(1) & (2) & (3) : 2.5
(1) & (2): 2.75
(1) & (2): 2.25
Baseline modelc
(1): p(wx + w2
y + [xy] + l) +BM (1): p(wx + wy + [xy]2) +BM
(2): p(wx + wy + [xy]) +BM
(2): p(wx + [xy]2) +BM
(3): p(wx + [xy]2) +BM
Error correction factorsd
(1): βw = 0.96, βr=1.00
(1): βw = 0.99, βr=1.36
(2): βw = 1.00, βr=1.10
(1): βw = 0.97, βr=1.65
(3): βw = 1.00, βr=1.85
(1): -
(2): -
(1): -
(2): -
Table A.9: Same as Table A.5 for targets 12-14. The HD 115617 photometry is affected by a strong systematic error that
prevented any searching for the transit of HD 115617 b.
22
M. Gillon et al.: Results of the Warm Spitzer transit search
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
HD 125612
60027
(1): 2010-04-09
(2): 2010-09-10
3.6
(1): 38110464
(2): 40313600
Spitzer pipeline version
S18.18.0
Exposure time [s]
NBCD
b
Duration [hr]
Photometric aperture [pixels]
Baseline modelc
Error correction factorsd
0.4
(1): 1790
(2): 1310
(1): 13.8
(2): 10.2
(1): 1.75
(1): p(w2
x + w2
(2): 1.9
y + [xy]3 + l) +BM
HD 134060
(1): 60027
(2): 90072
(1): 2010-04-13
(2): 2013-05-04
(1): 3.6
(2): 4.5
(1): 38110720
(S2): 46916096
(2): 46915840
(1): S18.18.0
(2): S19.1.0
(1): 0.02
(2): 0.1
(1): 6200
(2): 4480
(1): 6
(2): 11.5
(1): 1.75
(2): 2.25
HD 181433
90072
2012-12-03
4.5
(S): 46915584
(1): 46915328
(2): 46915072
S19.1.0
0.4
(1): 1563
(2): 1498
(1): 11.9
(2): 11.4
(1): 3.25
(2): 3.25
(2): p(t + wx + w2
y + [xy] + l2) +BM (2): p(w2
(1): p(wx + [xy]2 + l) +BM
(1): p(wx + wy + [xy]2) + BM
x + wy + [xy]4) +BM (2): p(wx + wy + [xy]2) + BM
(1): βw = 0.81, βr =2.02
(2): βw = 0.72, βr =2.34
(1): βw = 1.00, βr=1.62
(1): βw = 0.95, βr =1.27
(1): βw = 0.95, βr=1.30
(2): βw = 0.93, βr =1.11
Table A.10: Same as Table A.5 for targets 15-17.
Star
Program ID
Observation date
Channel [µm]
AOR(s)a
Spitzer pipeline version
Exposure time [s]
NBCD
b
HD 215497
90072
2010-07-16
4.5
38701568
S18.18.0
0.4
1750
Duration [hr]
13.4
Photometric aperture [pixels]
3
HD 219828
60027 & 90072
(1): 2010-08-01
(2): 2013-03-06
4.5
(1): 38702336
(2S): 46914304
(2A): 46914048
(2B): 46913792
S19.1.0
0.4
(1): 1440
(2A): 1564
(2B): 448
(1): 11
(2A): 11.9
(2B): 3.4
(1): 2.5
(2A): 3
(2B): 3
Baseline modelc
p(wx + wy + [xy] + l) +BM (1): p(wx + wy + [xy]2 + l2) +BM
(2A): p(wx + [xy]2 + l) +BM
(2B): p(wx + [xy]2 + l2) +BM
Error correction factorsd
βw = 0.95, βr=1.20
(1): βw = 0.93, βr=1.53
(2A): βw = 0.96, βr =1.01
(2B): βw = 0.94, βr=1.14
Table A.11: Same as Table A.5 for targets 18-19.
23
|
1301.7205 | 1 | 1301 | 2013-01-30T12:05:21 | The potential of sparse photometric data in asteroid shape modeling | [
"astro-ph.EP"
] | We investigate the potential of the sparse data produced by the Catalina Sky Survey astrometric project (CSS for short) in asteroid shape and rotational state determination by the lightcurve inversion method. We show that although the photometric quality of the CSS data, compared to the dense data, is significantly worse, it is in principle possible that these data are for some asteroids with high lightcurve amplitudes sufficient for a unique shape determination. CSS data are available for $\sim$180 asteroids for which shape models were previously derived from different photometric data sets. For 13 asteroids from this sample, we derive their unique shape models based only on CSS data, compare the two independent shape models together and discuss the reliability of models derived from only CSS data. We also use CSS data to determine shape models for asteroids with already known rotational period values, derive 12 unique models and compare previously published periods with periods determined from the full 3D modeling by the lightcurve inversion method. Finally, we test different shape resolutions used in the lightcurve inversion method in order to find reliable asteroid models. | astro-ph.EP | astro-ph | The potential of sparse photometric data in asteroid shape modeling
Astronomical Institute, Faculty of Mathematics and Physics, Charles University, V Holesovick´ach 2, 180 00 Prague, Czech Republic.
J. Hanus∗, J. Durech
3
1
0
2
n
a
J
0
3
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
0
2
7
.
1
0
3
1
:
v
i
X
r
a
Abstract
We investigate the potential of the sparse data produced by the Catalina Sky Survey astrometric project (CSS for short)
in asteroid shape and rotational state determination by the lightcurve inversion method. We show that although the
photometric quality of the CSS data, compared to the dense data, is significantly worse, it is in principle possible that
these data are for some asteroids with high lightcurve amplitudes sufficient for a unique shape determination. CSS
data are available for ∼ 180 asteroids for which shape models were previously derived from different photometric
data sets. For 13 asteroids from this sample, we derive their unique shape models based only on CSS data, compare
the two independent shape models together and discuss the reliability of models derived from only CSS data. We
also use CSS data to determine shape models for asteroids with already known rotational period values, derive 12
unique models and compare previously published periods with periods determined from the full 3D modeling by the
lightcurve inversion method. Finally, we test different shape resolutions used in the lightcurve inversion method in
order to find reliable asteroid models.
Keywords: Asteroids, Sparse photometry, Shape models, Lightcurves
1. Introduction
Orbital parameters are currently known for more than
400 000 asteroids. On the other hand, rotational states
and shapes were determined only for a small fraction
In the Minor Planet Lightcurve Database1
of them.
(Warner et al., 2009), periods for ∼ 3 500 asteroids
are stored. Convex shapes with spin axis directions
were derived only for ∼ 200 asteroids, these three-
dimensional models of asteroids are available in the
Database of Asteroid Models from Inversion Tech-
Durech et al., 2010) maintained by
niques2 (DAMIT,
the Astronomical Institute of the Charles University in
Prague, Czech Republic.
All asteroid models in DAMIT were derived by the
lightcurve inversion method (LI). This gradient-based
method is a powerful tool that allows us to derive ba-
sic physical properties of asteroids (the rotational state
and the shape) from their disk-integrated photometry
∗Corresponding author. Tel: +420 221912572.
Fax: +420
221912577.
Email address: [email protected] ()
1http://cfa-www.harvard.edu/iau/lists/Lightcurve-
Dat.html
2http://astro.troja.mff.cuni.cz/projects/aste-
roids3D
(see Kaasalainen and Torppa, 2001; Kaasalainen et al.,
2001, 2002). LI, in a form we use, leads to a convex
shape approximation of a real asteroid (close to its con-
vex hull) that has to be a single body (or the possible
moon has to be so small that its photometric signature is
indistinguishable from the noise) in a relaxed rotational
state (i.e., rotates around the axis with a maximum mo-
mentum of inertia). Sidereal rotational period and the
direction of the spin axis are determined simultaneously
with the shape. Typical number of free parameters is
∼50 -- 80.
We use two different types of disk-integrated photom-
etry: (i) dense-in-time, which typically consists of tens
to a few hundreds of individual data points observed
during one revolution (typically several hours), and (ii)
sparse-in-time, where the typical separation of individ-
ual measurements is large compared to the rotation pe-
riod. For sparse data, we usually have a few measure-
ments per night, such as in the case of astrometric sky
surveys. Based on the type of the photometry, we use
the terms "dense lightcurves" and "sparse lightcurves".
Unique asteroid shape models are typically feasible if
we have dense lightcurves from at least three apparition,
combined dense and sparse data give us more variabil-
ity.
Preprint submitted to Planetary and Space Science
September 2, 2018
Our long-term strategy is to enlarge the number of
asteroids with known shapes and rotational states, be-
cause it can help us in understanding the physical pro-
cesses that take place in the asteroid's populations,
such as Near Earth Asteroids (NEAs), Main Belt As-
teroids (MBAs) or even asteroids in individual fami-
lies (e.g., Koronis family, Slivan, 2002; Slivan et al.,
2003). Current distribution of periods and spin axes is
the direct result of the evolution of these objects start-
ing with their formation until the present time (several
hundreds of Myr to ∼ 4 Gyr for most studied aster-
oids). The most prominent processes acting on aster-
oids are collisions and mass shedding, and the YORP ef-
fect3 (Rubincam, 2000; Vokrouhlick´y et al., 2003). The
knowledge of the shape can be used for several pur-
poses, e.g.: (i) in the construction of a thermal model
(e.g., Muller et al., 2011), where values for geometric
albedo, size and surface properties can be determined,
(ii) a sample of real shapes instead of synthetic ones can
be used for the statistical study of the non-gravitational
forces (Yarkovsky4 and YORP effects), or (iii) in combi-
nation with star occultations by asteroids. These events
(observed for hundreds of asteroids) give us additional
information about the shape (e.g. non-convexities) and
can lead to a size estimation with a typical uncertainty
of 10% (see Durech et al., 2011).
Most of the currently available photometric data were
already used in the LI. The only significant excep-
tion are data from the Catalina Sky Survey astrometric
project (CSS for short, Larson et al., 2003). In this pa-
per, we investigate the possibility of determination of re-
liable asteroid shape models from only a small amount
of sparse-in-time distributed low-quality photometric
measurements (∼ 100) by the LI. Such data from CSS
project are available for several thousands of asteroids.
We want to find out if these data are for asteroids with
high lightcurve amplitudes of a sufficient amount and
quality for a unique shape determination, and if so, how
reliable these asteroid models are. The investigation
of sparse data capabilities and the reliability of derived
models is important, because (i) it can lead to a deter-
mination of new asteroid models without a need of ob-
serving any additional photometric lightcurves, and (ii)
in a few years, another huge amount of sparse data from
three astrometric surveys will be available -- from the
Pan-STARRS (Panoramic Survey Telescope and Rapid
3Yarkovsky -- O'Keefe -- Radzievskii -- Paddack
torque
caused by the recoil force from anisotropic thermal emission which
can alter rotational periods and orientation of spin axes
4a thermal force acting on a rotating asteroid which can alter semi-
effect,
a
major axes
2
Response System, Hodapp et al., 2004), and later also
from the Gaia satellite (Perryman et al., 2001), and the
LSST (Large Synoptic Survey Telescope, Ivezi´c et al.,
2008). Understanding the CSS data with respect to the
asteroid shape modeling will speed up future processing
and use of the new sparse data.
2. Photometry
For a unique5 shape determination of an asteroid,
we typically need dense photometric data from at least
three apparitions and a good coverage of the solar phase
angle6 (i.e., different viewing geometries), ∼ 20 such
dense lightcurves are usually sufficient. The biggest
limiting factor of these data is their insufficient amount.
Dense lightcurves were observed for ∼ 3 500 asteroids.
Only in ∼ 100 cases, asteroid shape models were de-
rived. Determination of new models based on the dense
data is now possible only if new lightcurves are ob-
served. Many amateurs or semi-professionals use their
telescopes for asteroid observations, but their main goal
is the determination of the synodic period, which can
be derived already from one apparition. If a period for
a particular asteroid is then secure, they usually do not
observe this object any more. We have to observe such
object by ourselves. This approach is demanding on ob-
servational time and also on fundings. Models based
only on dense data are important in testing the relia-
bility of models based on sparse data (due to possible
comparison).
Currently available sparse photometry is accessible
via the Asteroids -- Dynamic Site database7 (AstDyS),
where data from hundreds of astrometric observatories
are stored. The photometry is mostly a by-product of
astrometric measurements and in most cases, asteroid
magnitudes are given to only one decimal place, so the
accuracy is 0.1 mag at best. Whether or nor this is suffi-
cient for a unique shape determination for a reasonable
number of asteroids was studied in Hanus et al. (2011).
The authors have found 7 observatories with quality
data and used these data in combination with relative
lightcurves for asteroid shape modeling. The most ac-
curate sparse photometry is from the U.S. Naval Obser-
vatory in Flagstaff (USNO-Flagstaff station, MPC code
5We define a unique solution as follows: (i) the best period has at
least 10% lower χ2 than all other periods in the scanned interval, (ii)
for this period, there is only one pole solution with at least 10% lower
χ2 than the others (with a possible ambiguity for λ ± 180◦), and (iii)
this solution fulfills all our additional tests (discussed in Hanus et al.,
2011).
6the Sun -- asteroid -- Earth angle
7http://hamilton.dm.unipi.it/astdys/
689). These data were already studied in Durech et al.
(2009) and based on these data, the authors derived con-
vex shape models for 24 asteroids. The biggest dis-
advantage of the USNO-Flagstaff station data is that
they are available only for about 1000 brightest aster-
oids. Data from Hipparcos satellite (ESA, 1997) and
Roque de los Muchachos Observatory, La Palma (MPC
code 950) are also accurate (compared to the other
observatories), but their amount of measurements is
even smaller than for the USNO-Flagstaff station. The
largest amount of sparse photometry is available from
the Catalina Sky Survey observatory (CSS for short,
Larson et al., 2003). These data were already used in
the LI by Hanus et al. (2011), but only in combina-
tion with other photometric data. Their typical photo-
metric accuracy was ∼ 12%. Although this accuracy
of the CSS photometry seems very low, the data are
valuable for asteroids with lightcurve amplitudes higher
than & 0.3 mag, which is about a half of all aster-
oids (based on the lightcurve data of the Minor Planet
Lightcurve Database, Warner et al., 2009). There are
dense lightcurves for only ∼ 700 asteroids with num-
bers over ten thousand, on the other hand, we have at
least 100 sparse data points from CSS for ∼ 6 000 of
these asteroids.
3. Reliability of asteroid models derived only from
the Catalina Sky Survey photometric data
The Catalina Sky Survey astrometric project has pro-
duced the largest amount of sparse photometric mea-
surements. For most asteroids, no other data of suffi-
cient quality are available.
Test 1. One way how to test the reliability of mod-
els based only on CSS data is to compare them with
models derived from different photometric data set.
So far, about a hundred of asteroid models were de-
rived from dense data (e.g., Kaasalainen et al., 2002;
Torppa et al., 2003; Slivan et al., 2003; Marciniak et al.,
Durech et al., 2007), another ∼ 100 models
2007;
were derived from combined dense and sparse data
( Durech et al., 2009; Hanus et al., 2011). These models
are believed to be reliable and thus correct. For most
of these asteroids, CSS data are also available, so we
tried to derive their models from only CSS data itself.
We had at least 50 individual data points for 185 aster-
oids. In 13 cases, we were able to derive their unique
solutions. Only in 7 cases the solution was in an agree-
ment with the model based on a different data set, other
six solutions were formally correct, but derived periods
were different from the expected ones.
3
Test 2. The main inconvenient when computing mod-
els from sparse data is that synodic periods are known
only for ∼ 3 500 asteroids (determined from dense
lightcurves and stored in the Minor Planet Lightcurve
Database). This a priori information is used in the shape
modeling: we search for the model only near the as-
sumed period (e.g., we test the initial period values from
an interval (0.95 P, 1.05 P), where P is the published
period value). Typically, we test a huge number of ini-
tial period values (usually thousands), because the mod-
eled parameter space is full of local minims and the
gradient-based LI method converges to the nearest local
minimum corresponding to the initial period value. The
minimal difference between two local minims is given
mainly by the timespan of the observational data (see
Eq. 1) and is usually very small (. 10−3 h). Having a
short period interval saves a lot of computational time
and gives us confidence that the model is correct if the
derived period value is close to P. Unfortunately, for
most asteroids, we do not know the period (cannot be
directly determined from sparse data without the model
computation as in the case of the dense data), so we have
to search for the model on a period interval of all possi-
ble period values, typically 2 -- 100 hours. This consumes
a lot of computational time and we loose the possibility
of comparing the two period values.
In this test we used the a priori information about the
periods: for some asteroids, we knew their periods, we
had only CSS data (there exists their dense lightcurves,
but we do not have them in an electronic form), and,
based on that data, we were also able to derive their
models near the published period. We extended the
model search on the period interval of 2 -- 100 hours and
tried to find the model solutions again (as would be done
if the period was unknown). This was performed for
12 asteroids. In 7 cases, the previous solution derived
on a shorter period interval was reproduced also on a
larger period interval. For 4 asteroids, we did not get
a unique period solution on the extended period inter-
val, and a formally correct (i.e., it fulfilled our condi-
tions for a unique solution), but a different solution was
found for one asteroid. On Fig. 1, we show a peri-
odogram of asteroid (5647) 1990 TZ, where for each
initial period value a χ2-value corresponding to the best
shape model and pole direction (a local minimum in
the multi-dimensional parameter space) is plotted. This
model computation was based on 87 individual mea-
surements from CSS, derived period P = 6.13867 h is
in agreement with a period P = 6.141 h reported by
Bembrick and Bolt (2003).
2
1
0.8
0.6
0.4
0.2
0
(5547) 1990 TZ
Best solution, P = 6.13867 h
10
20
30
40
50
60
70
80
90
100
Period P [h]
Figure 1: Periodogram of asteroid (5647) 1990 TZ. Each point corresponds to a particular local minimum in the modeled parameter space of
periods, pole directions and shapes.
Test 3. Since we derived several unique models that
were clearly incorrect, we performed this test to check
more carefully the stability of the solution and to de-
tect false solutions. The shape of an asteroid is repre-
sented by coefficients of its expansion into the spherical
harmonic functions. These coefficients are optimized in
the LI. The number of coefficients is given by the order
n of the shape expansion we use, we call n the shape
resolution. In the LI, we typically use n = 6, which
corresponds to 49 coefficients ((n + 1)2). For asteroids
with . 100 sparse data points from CSS, so many co-
efficients could be to much and the LI could then lead
to a random unstable solution that is even unique and
physically correct. Another input parameter of the LI
is the period step Pstep between two subsequent periods
from the scanned interval. The minimal difference be-
tween two local minims in the modeled parameter space
of periods is given by
∆P =
P2
T
,
(1)
where P is the period value and T the timespan of the
observational data. In our computations, we use for the
period step a value Pstep = 0.8∆P.
We tested the stability of the solution on different val-
ues of the resolution n and on two values of the pe-
riod step Pstep (0.8∆P and 0.5∆P). Table 1 summa-
rizes the results of this test for all asteroids in Test 1
for which we derived their unique solutions for n = 6
and Pstep = 0.8∆P (corresponds to column 6). Correct
unique solution is marked by "X", wrong unique solu-
tion by "x" and no unique solution by " -- ".
4. Discussion
In Test 1, where we compared the derived models
with corresponding models based on different data sets
(stored in DAMIT), in 7 cases out of 13, the solution
was in a well agreement with a model based on a differ-
ent and higher data set. From this point of view, CSS
data, which represents the largest dataset of sparse pho-
tometry with photometric accuracy theoretically suf-
ficient for a unique shape determination, seem to be
promising for the asteroid shape modeling. Unfortu-
nately, several clearly wrong models that fulfilled our
conditions on a unique solution were derived (they had
different sidereal rotational periods than was expected).
Similar problem occurred in Test 2, where we compared
the derived models with their previously known periods.
One model was clearly incorrect.
There are several possible explanations why we de-
rived incorrect models from the CSS data: (i) data for
asteroids with low amplitudes are more noisy, (ii) low
amount of data with respect to the number of modeled
parameters, (iii) systematic errors of the survey. In cases
(i) and (ii), the data are poor, contain not enough infor-
mation about the period and shape, and could produce a
random solution with an unrealistic rms value (the fit
is "too perfect"). Oszkiewicz et al. (2011) show that
photometric data sets from astrometric stations (such
as CSS) exhibit magnitude variations with apparent V-
band magnitude. For the brightest asteroids, the images
are saturated, and for the faintest, the background sub-
traction is imperfect. This consequence of (iii) is prob-
ably partially visible in our data -- most of the incorrect
4
c
Table 1: Test of the stability of the model solution on the shape resolution n and the step in the period Pstep. First column gives the asteroid number,
symbol "X" means that for a given n and Pstep, both unique and correct solution was derived, symbol "x" means that a unique but wrong solution
was derived, and finally, in the case if no unique solution was determined we use the symbol " -- ".
Pstep = 0.8∆P
Pstep = 0.5∆P
n=2
Ast
685
X
1022 X
1419 X
1568 X
2156 X
--
3678
4483 X
--
52
68
--
216
--
--
455
--
984
--
1382
n=3
X
X
X
X
X
--
X
--
--
--
--
--
--
n=4
--
--
X
X
X
X
X
x
--
--
x
--
--
n=5
X
--
X
X
X
X
X
--
--
x
x
--
--
n=6
X
X
X
X
X
X
X
x
x
x
x
x
x
n=2
X
X
X
X
X
--
X
--
--
--
--
--
--
n=3
X
X
X
X
X
--
X
--
--
--
--
--
--
n=4
n=5
--
--
X
X
X
X
X
x
--
--
x
--
--
--
--
X
X
X
--
X
--
--
--
x
x
--
n=6
X
X
X
X
X
X
X
--
x
--
x
--
--
models are brighter (have higher absolute magnitudes)
than the correct ones. We are not able to distinguish be-
tween the effects of the amplitude and the noise from
each other only from the sparse data itself without a
model computation, so there is no way how to detect
the case (i). The amount of data sufficient for a cor-
rect unique shape determination is dependent mainly on
the time epochs (geometry of observations) and also on
the signal-to-noise ratio, which we do not know. So,
the sufficient amount of sparse measurements strongly
varies from one asteroid to the other.
As seen from Table 1, models derived correctly for
resolution n = 6 and period step Pstep = 0.8∆P are
mostly determined for other resolutions n and both pe-
riod steps Pstep. No wrong solution appeared. In several
cases, a unique solution was not found, but the correct
period was still detected (we had multiple pole solu-
tions). On the other hand, for incorrect models from
Test 1, we got unique solutions for different resolutions
n and period steps Pstep only rarely and never for n = 2
or n = 3. All these unique solutions were always also
incorrect. There is a weak dependence on the period
step, Pstep = 0.5∆P seems to be more convenient be-
cause when used, less incorrect models were derived.
To minimize the determination of incorrect models that
satisfy our requirement on unique solutions, we should
use the values n = 6 and Pstep = 0.5∆P for resolution
and period step, respectively. If a unique solution is then
determined, the computation should be rerun with n = 3
and Pstep = 0.5∆P. The model is then reliable if we get
similar unique solutions.
We rerun the computation for all 8 successfully de-
rived models from Test 2 with n = 3 and Pstep = 0.5∆P.
Previously correct solutions were confirmed, the incor-
rect one was ruled out.
Our method assumes a single body in a relaxed ro-
tation state (rotates around the principal axis with the
maximum momentum of inertia). Objects like tumblers
or binary asteroids are not identified by the used form
of the LI, their model determination usually fails be-
cause the single period model does not cover the more
complex period features. The only exception are fully
synchronous close binary systems with similar-sized
components, such as 91 Antiope (Merline et al., 2000).
These objects are reproduced as highly elongated single
bodies, and because they have typically high lightcurve
amplitudes, they could be easily present in the possible
sample of models based on CSS data.
5. Conclusions
Sparse data from the Catalina Sky Survey are low ef-
ficient in unique model determination and seem more
accurate for less brighter asteroids. Incorrectly deter-
mined models could be indicated by unrealistic rms val-
ues, but not generally.
Correct period solutions for a single asteroid but dif-
ferent resolution n were always stable, the dispersion in
the period was typically lower than ∆P (given by Eq. 1),
5
which is for a period value of 10 hours and the timespan
of the observations 10 years ∼ 10−3 h (typical values
for the period and the timespan of the CSS data). The
dispersion in the ecliptic latitude of the pole direction
was usually . 20◦, but in the ecliptic longitude of the
pole direction often & 50◦. This significant dispersion
in longitude was partially caused by models with high
values of ecliptic latitudes (longitudes are more dense
near the pole than near the equator) and because of the
small resolution of the shapes.
We showed that CSS data can be successfully used
for reliable asteroid model determination. The most ac-
curately derived parameter is the rotational period. Al-
though the Gaia satellite will produce during its oper-
ational time in average only ∼ 60 data points for each
asteroid (for CSS, we have ∼ 100), the photometric ac-
curacy will be incomparably better than for CSS data.
Less data will be compensated by better accuracy al-
lowing determination of many new asteroid models (or
at least accurate values for rotational periods). The com-
bination of Gaia data with data from Pan-STARRS and
LSST projects will significantly increase the amount of
new models. For asteroids with low amplitudes, their
signal-to-noise ratio of the Gaia data will be also low
and we will encounter the same problem of incorrect
solutions as for CSS data. The detection of these false
models will be similar.
Finding an infallible test for detecting the false so-
lutions derived from Catalina data could allow us de-
termination of shapes and physical parameters for tens
of asteroids. While the LI is more successful for as-
teroids with higher lightcurve amplitudes, the derived
sample of asteroid parameters will be significantly bi-
ased. For example, shapes will be more elongated than
it is usual in the whole population. Careful de-biasing
of the model sample will be necessary before its use in
any statistical study and physical interpretation.
Possible models based on Catalina data are good can-
didates for follow-up observations, new dense photom-
etry should help in the detection of incorrect solutions
and should lead to more detailed shape models.
Acknowledgements
The work of JH and JD has been supported by grant
GA UK 134710 of the Grant agency of the Charles Uni-
versity, by the project SVV 265301 of the Charles Uni-
versity in Prague, by grants GACR 205/08/H005 and
GACR 209/10/0537 of the Czech grant agency and by
the Research Program MSM0021620860 of the Czech
Ministry of Education. The calculations were per-
formed on the computational cluster Tiger at the As-
tronomical Institute of the Charles University in Prague
(http://sirrah.troja.mff.cuni.cz/tiger).
References
Bembrick, C., Bolt, G., 2003. Lightcurves across Australia: period
determination for minor planet (5647) 1990 TZ. Minor Planet Bul-
letin 30, 42 -- 43.
Durech, J., Kaasalainen, M., Herald, D., Dunham, D., Timerson, B.,
Hanus, J., Frappa, E., Talbot, J., Hayamizu, T., Warner, B.D.,
Pilcher, F., Gal´ad, A., 2011. Combining asteroid models derived by
lightcurve inversion with asteroidal occultation silhouettes. Icarus
214, 652 -- 670. 1104.4227.
Durech, J., Kaasalainen, M., Warner, B.D., Fauerbach, M., Marks,
S.A., Fauvaud, S., Fauvaud, M., Vugnon,
J., Pilcher, F.,
Bernasconi, L., Behrend, R., 2009. Asteroid models from com-
bined sparse and dense photometric data. Astron. Astrophys. 493,
291 -- 297.
Durech, J., Scheirich, P., Kaasalainen, M., Grav, T., Jedicke, R., Den-
neau, L., 2007. Physical models of asteroids from sparse photo-
metric data, in: G. B. Valsecchi, D. Vokrouhlick´y, & A. Milani
(Ed.), IAU Symposium, pp. 191 -- 200.
Durech, J., Sidorin, V., Kaasalainen, M., 2010. DAMIT: a database of
asteroid models. Astron. Astrophys. 513, A46.
ESA, 1997. The Hipparcos and Tycho Catalogues (ESA 1997).
VizieR Online Data Catalog 1239, 0.
Hanus, J., Durech, J., Broz, M., Warner, B.D., Pilcher, F., Stephens,
R., Oey, J., Bernasconi, L., Casulli, S., Behrend, R., Polishook,
D., Henych, T., Lehk´y, M., Yoshida, F., Ito, T., 2011. A study
of asteroid pole-latitude distribution based on an extended set of
shape models derived by the lightcurve inversion method. Astron.
Astrophys. 530, A134. 1104.4114.
Hodapp, K.W., Kaiser, N., Aussel, H., Burgett, W., Chambers, K.C.,
Chun, M., Dombeck, T., Douglas, A., Hafner, D., Heasley, J.,
Hoblitt, J., Hude, C., et al., 2004. Design of the Pan-STARRS
telescopes. Astronomische Nachrichten 325, 636 -- 642.
Ivezi´c, Z., Tyson, J.A., Acosta, E., Allsman, R., Anderson, S.F., An-
drew, J., Angel, R., Axelrod, T., Barr, J.D., Becker, A.C., Be-
cla, J., Beldica, C., et al., 2008. LSST: from Science Drivers to
Reference Design and Anticipated Data Products. ArXiv e-prints
0805.2366.
Kaasalainen, M., Mottola, S., Fulchignoni, M., 2002. Asteroid Mod-
els from Disk-integrated Data. Asteroids III , 139 -- 150.
Kaasalainen, M., Torppa, J., 2001. Optimization Methods for Asteroid
Lightcurve Inversion. I. Shape Determination. Icarus 153, 24 -- 36.
Kaasalainen, M., Torppa, J., Muinonen, K., 2001. Optimization Meth-
ods for Asteroid Lightcurve Inversion. II. The Complete Inverse
Problem. Icarus 153, 37 -- 51.
Larson, S., Beshore, E., Hill, R., Christensen, E., McLean, D., Kolar,
S., McNaught, R., Garradd, G., 2003. The CSS and SSS NEO sur-
veys, in: AAS/Division for Planetary Sciences Meeting Abstracts
#35, p. 982.
Marciniak, A., Michałowski, T., Kaasalainen, M.,
Durech, J.,
Poli´nska, M., Kwiatkowski, T., Kryszczy´nska, A., Hirsch, R.,
Kami´nski, K., Fagas, M., Colas, F., Fauvaud, S., Santacana, G.,
Behrend, R., Roy, R., 2007. Photometry and models of selected
main belt asteroids. IV. 184 Dejopeja, 276 Adelheid, 556 Phyllis.
Astron. Astrophys. 473, 633 -- 639.
Merline, W.J., Close, L.M., Dumas, C., Shelton, J.C., Menard, F.,
Chapman, C.R., Slater, D.C., 2000. Discovery of Companions
to Asteroids 762 Pulcova and 90 Antiope by Direct Imaging, in:
AAS/Division for Planetary Sciences Meeting Abstracts #32, p.
1017.
6
Muller, T.G., Durech, J., Hasegawa, S., Abe, M., Kawakami, K., Ka-
suga, T., Kinoshita, D., Kuroda, D., Urakawa, S., Okumura, S.,
Sarugaku, Y., Miyasaka, S., Takagi, Y., Weissman, P.R., Choi, Y.J.,
Larson, S., Yanagisawa, K., Nagayama, S., 2011. Thermo-physical
properties of 162173 (1999 JU3), a potential flyby and rendezvous
target for interplanetary missions. Astron. Astrophys. 525, A145.
1011.5029.
Oszkiewicz, D.A., Muinonen, K., Bowell, E., Trilling, D., Penttila, A.,
Pieniluoma, T., Wasserman, L.H., Enga, M.T., 2011. Online multi-
parameter phase-curve fitting and application to a large corpus of
asteroid photometric data. Journal of Quantitative Spectroscopy &
Radiative Transfer 112, 1919 -- 1929.
Perryman, M.A.C., de Boer, K.S., Gilmore, G., Høg, E., Lat-
tanzi, M.G., Lindegren, L., Luri, X., Mignard, F., Pace, O.,
de Zeeuw, P.T., 2001. GAIA: Composition,
formation and
evolution of the Galaxy. Astron. Astrophys. 369, 339 -- 363.
ArXiv:astro-ph/0101235.
Rubincam, D.P., 2000. Radiative Spin-up and Spin-down of Small
Asteroids. Icarus 148, 2 -- 11.
Slivan, S.M., 2002. Spin vector alignment of Koronis family asteroids.
Nature 419, 49 -- 51.
Slivan, S.M., Binzel, R.P., Crespo da Silva, L.D., Kaasalainen, M.,
Lyndaker, M.M., Krco, M., 2003. Spin vectors in the Koronis
family: comprehensive results from two independent analyses of
213 rotation lightcurves. Icarus 162, 285 -- 307.
Torppa, J., Kaasalainen, M., Michałowski, T., Kwiatkowski, T.,
Kryszczy´nska, A., Denchev, P., Kowalski, R., 2003.
Shapes
and rotational properties of thirty asteroids from photometric data.
Icarus 164, 346 -- 383.
Vokrouhlick´y, D., Nesvorn´y, D., Bottke, W.F., 2003. The vector align-
ments of asteroid spins by thermal torques. Nature 425, 147 -- 151.
Warner, B.D., Harris, A.W., Pravec, P., 2009. The asteroid lightcurve
database. Icarus 202, 134 -- 146.
7
|
1703.05338 | 1 | 1703 | 2017-03-15T18:03:19 | On the origin of the spiral morphology in the Elias 2-27 circumstellar disc | [
"astro-ph.EP",
"astro-ph.SR"
] | The young star Elias 2-27 has recently been observed to posses a massive circumstellar disc with two prominent large-scale spiral arms. In this Letter we perform three-dimensional Smoothed Particle Hydrodynamics simulations, radiative transfer modelling, synthetic ALMA imaging and an unsharped masking technique to explore three possibilities for the origin of the observed structures -- an undetected companion either internal or external to the spirals, and a self-gravitating disc. We find that a gravitationally unstable disc and a disc with an external companion can produce morphology that is consistent with the observations. In addition, for the latter, we find that the companion could be a relatively massive planetary mass companion (less than approximately 10 - 13 MJup) and located at large radial distances (between approximately 300 - 700 au). We therefore suggest that Elias 2-27 may be one of the first detections of a disc undergoing gravitational instabilities, or a disc that has recently undergone fragmentation to produce a massive companion. | astro-ph.EP | astro-ph |
Draft version September 4, 2018
Preprint typeset using LATEX style emulateapj v. 08/22/09
ON THE ORIGIN OF THE SPIRAL MORPHOLOGY IN THE ELIAS 2 -- 27 CIRCUMSTELLAR DISC
Farzana Meru, Attila Juh´asz, John D. Ilee, Cathie J. Clarke, Giovanni P. Rosotti, and Richard A. Booth
Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge, CB3 0HA, UK
Draft version September 4, 2018
ABSTRACT
The young star Elias 2-27 has recently been observed to posses a massive circumstellar disc with
two prominent large-scale spiral arms. In this Letter we perform three-dimensional Smoothed Particle
Hydrodynamics simulations, radiative transfer modelling, synthetic ALMA imaging and an unsharped
masking technique to explore three possibilities for the origin of the observed structures -- an unde-
tected companion either internal or external to the spirals, and a self-gravitating disc. We find that
a gravitationally unstable disc and a disc with an external companion can produce morphology that
is consistent with the observations. In addition, for the latter, we find that the companion could be
a relatively massive planetary mass companion ((cid:46) 10 − 13MJup) and located at large radial distances
(between ≈ 300 − 700 au). We therefore suggest that Elias 2-27 may be one of the first detections of
a disc undergoing gravitational instabilities, or a disc that has recently undergone fragmentation to
produce a massive companion.
Subject headings: stars:
individual (Elias 2-27) -- stars: pre-main sequence -- hydrodynamics --
radiative transfer -- protoplanetary disks -- planet-disk interactions
1.
INTRODUCTION
With the advent of the Atacama Large Millimetre Ar-
ray (ALMA), it has for the first time become possible
to spatially resolve, and thus directly observe, the mid-
plane structure of protoplanetary discs where planet for-
mation processes occur. Such an extreme increase in ob-
servational capability has given rise to several surprising
results, examples of which include the symmetric ring
structures in HL Tau (ALMA Partnership et al. 2015),
TW Hydrae (Andrews et al. 2016) and HD 163296 (Isella
et al. 2016), the horseshoe shaped dust traps in HD
142527 (Casassus et al. 2013), and the birth of a ternary
system via disc fragmentation in L1448 IRS3B (Tobin
et al. 2016). Observations such as these are extremely
powerful, as sub-structure within protoplanetary discs
can be a signpost of dynamical or chemical effects occur-
ring within the star-disc system. Therefore, a proper
understanding of their cause is essential to determine
which of these processes are important during the star
and planet formation process.
A recent example of such a spatially-resolved observa-
tion was presented by P´erez et al. (2016) in which Elias
2-27 was targeted with ALMA. Elias 2-27 is a low mass
young star (M(cid:63) = 0.5 -- 0.6 M(cid:12), tage ∼ 1 Myr; Luhman
& Rieke 1999; Natta et al. 2006). Based on its spectral
energy distribution (SED), the system is thought to be-
long to the Class II phase (Andrews et al. 2009; Evans
et al. 2009), yet at the same time, observations have sug-
gested an unusually large disc mass, ranging from 0.04
-- 0.14 M(cid:12) (Andrews et al. 2009; Isella et al. 2009; Ricci
et al. 2010).
The Elias 2-27 disc posseses two large-scale symmet-
ric spiral arms (Figure 1, right). Additionally when the
raw ALMA observations are processed with an unsharp
masking filter, two dark crescents interior to the spirals
Electronic address: [email protected]
Electronic address: [email protected]
Electronic address: [email protected]
and a bright inner ellipse are revealed (Figure 1, left).
The origin of these is unclear. With such a large disc-to-
star mass ratio, could the disc be self-gravitating? Or,
could an as-yet-undetected companion be causing these
features via dynamical interactions?
In this Letter we describe the results of hydrodynam-
ical and radiative transfer modelling of the Elias 2-27
system. We produce synthetic ALMA observations to
explore three possibilities that may give rise to the ob-
served features -- a companion internal to the spirals, a
companion external to the spirals, or gravitational insta-
bilities operating within the disc.
2. METHODOLOGY
2.1. Hydrodynamics
Our
simulations are performed using a three-
dimensional Smoothed Particle Hydrodynamics code
(sph) which includes the heating due to work done and
the radiative transfer of energy in the flux-limited diffu-
sion limit (Whitehouse et al. 2005; Whitehouse & Bate
2006). A detailed description of the code can be found in
Meru (2015), however for this work we employ two dif-
ferences. Firstly, boundary particles are located at every
timestep, allowing the vertical location between the op-
tically thick and thin region to be regularly re-evaluated,
leading to more accurate boundary temperatures. Sec-
ondly, we employ the Morris & Monaghan (1997) arti-
ficial viscosity with the sph parameter, αSPH, varying
between 0.1 -- 1.0 and βSPH = 2αSPH, to model shocks
within the disc.
We perform 72 hydrodynamical simulations varying a
number of disc properties as well as orbit properties for
the companion simulations. We first describe our refer-
ence disc setup: we model a 0.5M(cid:12) star surrounded by
a disc whose temperature follows
(cid:18) R
(cid:19)−q
200 au
T (R) = 13.4K
.
(1)
2
Meru et al.
Figure 1. Left: 1.3 mm continuum image of Elias 2-27, processed with an unsharp masking filter (originally presented by P´erez et al.
2016). Two symmetric spiral arms, a bright inner ellipse, and two dark crescents are clearly visible. The beam is shown in the lower left
corner as a filled ellipse. Right: Deprojection of the original 1.3mm image with an r2 scaling applied, showing two prominent spiral arms.
The white star denotes the central star's location. The apparent ring structure and the central ellipse in the left image are artefacts of the
unsharp masking.
The disc surface density follows
(cid:18) R
(cid:19)−p
Rc
(cid:32)
exp
−
(cid:19)2−p(cid:33)
(cid:18) R
Rc
Σ(R) = Σc
,
(2)
where Σc is the surface mass density of gas at the cut-off
radius, Rc. We model this disc between Rin = 10 au
and Rout = 400 au using the parameters determined by
P´erez et al. (2016), namely Σc = 5 g cm−2, Rc = 200 au,
q = 0.45 and p = 0.7.
We then perform a suite of simulations with a wide
range of parameters to test the internal companion, ex-
ternal companion and gravitational instability hypothe-
ses. Due to the computational expense of each simula-
tion, our aim is not to fit the exact parameters, but to
test whether each hypothesis can reproduce the observed
morphology. We vary the disc mass, surface mass den-
sity profile, temperature profile, cut-off radius, and outer
disc radius. We also model pure power surface density
profiles, i.e. without the exponential term in equation 2,
so that, once evolved, the disc has a much steeper pro-
file in the outer regions. In addition, for the companion
simulations, we vary the companion mass, the pericentre
distance, eccentricity and inclination.
Each disc is modelled using 250,000 sph gas particles.
The ratio of the smoothing length to the disc scale-height
is < 0.5 outside 40 -- 60 au, giving sufficient resolu-
tion to probe dynamical effects on the disc. The discs
are modelled with radiation hydrodynamics using flux-
limited diffusion, and the surface temperature (represent-
ing irradiation from the central star) is held at the profile
given by Eqn. 1. For the parameters studied here, the
disc is optically thick to stellar irradiation, and remains
vertically isothermal at the boundary temperature be-
yond a radius of 20 and 30 au in the companion and
self-gravitating disc simulations, respectively. This im-
plies that the thermodynamics are mainly set by external
irradiation.
We assume that the gas and dust are well mixed. For
our self-gravitating simulations the Stokes numbers of
millimeter particles are O(0.01), and thus dust trapping
in the spirals is expected to be marginal (Booth & Clarke
2016; Shi & Chiang 2014).
In our simulations with a
companion the Stokes numbers are O(0.1). For a spi-
ral to trap dust, however, it is also necessary that the
timescale for it to concentrate towards the pressure max-
imum (Clarke & Lodato 2009) is less than the crossing
timescale of a spiral feature. Although this condition is
readily met in the case of self-gravitating discs where the
spirals nearly co-rotate with the Keplerian flow, this is
not true in the case of a planet generated spiral which
co-rotates with the planet (e.g. Paardekooper & Mellema
2006; Zhu et al. 2012; Birnstiel et al. 2013)
2.2. Radiative Transfer & Synthetic Imaging
To calculate synthetic observations of our models, we
use the 3D radiative transfer code radmc-3d1. The dust
opacity is calculated from the optical constants for as-
tronomical silicates (Weingartner & Draine 2001) using
Mie-theory, assuming a power-law grain size distribution
between 0.005 -- 1000 µm with a power exponent of −3.5.
Outside 80 au we obtain the dust temperature and den-
sity on the spherical mesh by interpolating the gas den-
sity and temperature from the hydrodynamic simulations
using an sph interpolation with a cubic spline smoothing
kernel (Monaghan 1992) and assume a uniform gas-to-
dust ratio of 100. Due to the accretion of sph particles
on the central star, we extrapolate the density inwards of
80 au using a radial power-law with an exponent chosen
to give a smooth transition from the results of the hy-
drodynamic modelling. We assume a vertical Gaussian
density distribution, whose scaleheight is calculated from
the temperature given by Eq. 1. The dust and gas tem-
peratures are assumed to be identical. Using these tem-
peratures and densities on a spherical mesh, we calculate
images at λ = 1.3 mm using the raytracer in radmc-3d.
1 http://www.ita.uni-heidelberg.de/dullemond/software/radmc-
3d/
'ec.ALMA6SirDl Drms(lliSseCrescents16h26m45.1s45.0s44.9s-24°23'07.0"08.0"09.0"10.0"5.A.10-210-1Sν (DrEitrDry units)-300-200-1000100200300[au]-300-200-1000100200300[au]1.01.21.51.82.02.22.52.83.0r2Sν (arbitrary units)On the disc morphology in Elias 2 -- 27
3
Figure 2. Schematic diagram of the Elias 2-27 system based on the results of P´erez et al. (2016). Also shown are the companion mass
limits based on observations (i) by Ratzka et al. (2005), (ii) from the UKIDSS data and (iii) theoretical gap-opening limits (see Section
2.3.2).
Synthetic observations for ALMA are calculated us-
ing the Common Astronomy Software Application v4.5
(CASA, McMullin et al. 2007). Visibilities are calculated
with the simobserve task while imaging is performed
with the clean task. Since our goal is not to fit the ob-
servations exactly, but merely to show the morphological
similarities between our models and the observations, we
do not use exactly the same (u, v) co-ordinates as the ob-
servations to calculate our visibilities. Instead, we choose
an antenna configuration in CASA (alma.out13) which
results in a very similar synthesised beam to that of the
real data: 0.27(cid:48)(cid:48)×0.25(cid:48)(cid:48) with PA = 86◦. To ensure our
synthetic observations are as close to the real data as
possible, we use a bandwidth of 6.8 GHz, assume a pre-
cipitable water vapour column of 2.7 mm, and a total
integration time of 725 seconds.
We then apply an unsharp masking filter to the images
in a similar manner to P´erez et al. (2016). This involves
convolving the image with a Gaussian kernel with a full
width half maximum (FWHM) of 0.33(cid:48)(cid:48) and subtracting
a scaled version of the result from the original image.
This acts to remove large scale emission and boost the
contrast of small scale structures.
2.3. Observational constraints
We describe various observational constraints that ap-
ply to the Elias 2-27 system, which we use when testing
the three possible scenarios.
2.3.1. Constraints on the total system mass
The 12CO J = 2 − 1 emission toward Elias 2-27 seems
to be undergoing Keplerian rotation out to ∼ 630 au,
and according to kinematic modelling, the enclosed mass
interior to the emission is 0.5±0.2M(cid:12) (P´erez et al. 2016).
However, absorption by the surrounding molecular cloud
significantly obscures the red-shifted component of the
emission, leading to some uncertainty in these derived
masses.
2.3.2. Constraints on the mass of a potential companion
A volume-limited multiplicity survey of the ρ-Ophiuci
molecular cloud was performed by Ratzka et al. (2005).
Elias 2-27 was found to be a single star with an up-
per limit on the K-band contrast of 2.5 mag between
0.13 -- 6.4(cid:48)(cid:48). Additionally, the area around Elias 2-27 has
been targeted by the UKIRT Infrared Deep Sky Sur-
vey (UKIDSS, e.g. Lawrence et al. 2007). The closest
point source is located approximately 14(cid:48)(cid:48) to the South
(16h26m44.95s, −24◦23(cid:48)21.88(cid:48)(cid:48)) with a K-band magni-
tude of 15.9 mag, giving a K-band contrast of 7.5 mag
with Elias 2-27. The limiting magnitude of the survey
in the K-band (17.8 mag) suggests a maximum contrast
with Elias 2-27 of 9.4 mag for any undetected sources.
Using the NextGen atmosphere models (Allard et al.
1997; Baraffe et al. 1998; Hauschildt et al. 1999), we con-
vert these K-band contrasts to upper limits for the mass
of any potential companion, Mcomp, for various orbital
distances from Elias 2-27. Assuming all objects lie at
the same distance as Elias 2-27 (139 pc), the UKIDSS
data excludes any unseen companions > 0.01M(cid:12) beyond
420au (with the exception of the closest source mentioned
above, which would translate to Mcomp ≈ 0.02M(cid:12) at
2000 au). Inside 420au a companion up to 0.08M(cid:12) could
be present based on Ratzka et al. (2005).
P´erez et al. (2016) present 12CO, C18O and 13CO ob-
servations which combined, show the presence of gas out
to ≈ 630 au. Based on their channel maps and the PV
diagram of 12CO, there is no strong indication of a gap
in the gas (though we note that the signal-to-noise is
low). In order to not open up a gap in the gas a planet
must satisfy the viscosity and pressure conditions for gap-
opening (Equations 68 and 69 of Lin & Papaloizou 1993;
see also Crida et al. 2006). Between 300 − 420 au (i.e.
4
Meru et al.
Figure 3. Top: Unsharp masked images of a disc with an internal companion (left), external companion (middle) and a gravitationally
unstable disc (right). The internal companion is not compatible with the observations. The external companion simulations and the
gravitationally unstable disc simulations show the closest match to the observation in Figure 1. Bottom: Deprojected mock ALMA images
with an r2 scaling of the simulations in the top panel (analogous to Figure 1 right). The intensity of the gravitationally unstable disc has
been scaled down by a factor of 1.5 before the r2 scaling is applied (see Section 4). The simulations are run for 1.6, 1.2 and 3.75 orbits at
350au in the internal companion, external companion and self-gravitating disc simulations, respectively.
from the spirals to where the UKIDSS observations be-
come relevant), the pressure condition is more stringent
for α (cid:46) 0.015 providing a gap-opening mass of ≈ 13MJup
(though this mass limit is likely to be higher for migrating
planets; Malik et al. 2015). Beyond 420au the UKIDSS
limit is more stringent than any realistic gap-opening
mass. Figure 2 shows a schematic diagram illustrating
the observational and theoretical constraints.
3. RESULTS
We require the unsharp masked synthetic observations
of the simulated discs to display morphology that is con-
sistent with the observations. This constitutes three
main features -- i) two large-scale symmetric spiral arms,
ii) two dark crescents interior to the spirals, and iii) a
bright inner ellipse along the major axis of the disc (see
Figure 1, left). The spiral arms are visible in the un-
sharped masking and deprojected images while the dark
crescents and bright inner ellipse are only present in the
unsharped masking image. The discs presented in Fig-
ure 3 employ a steep outer disc edge rather than an ex-
ponentially tapered disc.
3.1. Internal companion
Figure 3 (left) shows the simulated observation for one
of our internal companion simulations. The disc is a
0.08M(cid:12) disc with Σ ∝ R−0.75 and T ∝ R−0.75, and in-
cludes a 0.01M(cid:12) companion at 140 au that is allowed to
accrete from the disc and grow to approximately 0.03M(cid:12).
The companion clears a large gap in the disc, forming the
required central elliptical feature, but without the dark
crescents or two armed spiral as in the original unsharped
masking observations. A lower companion mass does not
generate the large-scale spirals, while higher mass com-
panions remove large amounts of material from the disc.
We therefore suggest that the morphology in Elias 2-27 is
unlikely to be due to an undetected companion internal
to the spirals.
3.2. External companion
Figure 3 (centre) shows the simulated observations for
one of our external companion simulations. This disc is
the same as that in Section 3.1 but includes a ≈ 10MJup
companion at ≈ 425 au. The simulated unsharped mask-
ing observation reproduces the large-scale spiral arms,
the dark crescents and the bright inner ellipse, while the
deprojected image shows two large-scale spirals analo-
gous to Figure 1 (right). We also note that simula-
tions with companions located much beyond the gas disc
((cid:38) 700 au) can only reproduce the observed morphology
by violating the companion mass limits in Section 2.3.2.
3.3. Gravitationally unstable disc
EXTERNAL16h26m45.1s45.0s44.9s-24°23'06.0"07.0"08.0"09.0"10.0"5.A.'ec.10-210-1Sν (DrbitrDry units)G.I.16h26m45.1s45.0s44.9s-24°23'06.0"07.0"08.0"09.0"10.0"5.A.'ec.10-210-1Sν (DrbitrDry units)INTERNAL16h26m45.1s45.0s44.9s-24°23'06.0"07.0"08.0"09.0"10.0"5.A.'ec.10-210-1Sν (DrbitrDry units)-300-200-1000100200300[au]-300-200-1000100200300[au]2.02.22.52.83.03.23.53.8r2Sν (arbitrary units)-300-200-1000100200300[au]-300-200-1000100200300[au]2.02.22.52.83.03.23.53.8r2Sν (arbitrary units)-300-200-1000100200300[au]-300-200-1000100200300[au]1.01.21.51.82.02.22.52.83.0r2Sν (arbitrary units)On the disc morphology in Elias 2 -- 27
5
Figure 3 (right) shows the simulated observation for
a gravitationally unstable disc with Σ ∝ R−0.5 and
T ∝ R−0.75. The disc mass is 0.24 M(cid:12), leading to a disc-
to-star mass ratio of 0.49. Even with this relatively mas-
sive disc, the combined system mass lies within the limits
discussed in Section 2.3.1. The simulated unsharp mask-
ing observation reproduces the large-scale spiral arms,
the dark crescents and the bright inner ellipse, while the
deprojected image shows two large-scale spirals analo-
gous to Figure 1 (right).
4. DISCUSSION
The combination of our hydrodynamic modelling and
simulated observations allows us to put strong con-
straints on the disc structure from which the sub-
millimetre continuum emission originates. Throughout
all of our models, we find that a steeply declining surface
mass density beyond ≈ 300 au is key to producing a close
match to the masked image. Figure 4 shows the scaled
radial intensity profile of our successful self-gravitating
disc and the radial intensity profile of the successful com-
panion simulations compared to the observational data
from P´erez et al. (2016) and the model of Andrews et al.
(2009). The steep decline beyond ≈ 300 au in our sim-
ulation manifests itself in the form of a steep emission
profile at large radii that turns over at approximately
the right radius, matching the observed profile well. The
modest scaling factor for the self-gravitating disc (∼ 1.5)
is well within uncertainties associated with dust-to-gas
ratios and/or grain opacities for circumstellar discs.
We note that
the resulting Toomre profile dips
marginally below 1 for a limited radial range in our suc-
cessful self-gravitating disc. However, this disc shows no
sign of fragmentation even though we evolve it for many
dynamical times for this radial range. On the other hand
a companion beyond the spirals ((cid:38) 300 au) is unlikely
to have formed by core accretion. This suggests that,
should the companion hypothesis be correct, this would
hint that the Elias 2-27 disc may have formed a fragment
by gravitational instability in the past.
We also stress the importance of applying an unsharp
mask to our simulated observations. While such a mask
is primarily applied to increase the contrast of the image,
subtle changes are introduced to the resulting image that
allow us to remove models from consideration. Changes
in the surface density profile affect the radial locations
at which there is an excess or deficit with respect to a
Gaussian mask. This is illustrated in Figure 5 which
shows a disc modelled using the properties presented by
P´erez et al. (2016) (see Section 2.1 for details) i.e. with an
exponential surface mass density profile, which does not
reproduce the observations well. Therefore while a range
of models replicate the spirals seen in the unprocessed
ALMA image (e.g. Tomida et al. 2017), comparison of
masked images constrains the disc properties further.
However, special care must be taken when interpreting
unsharp masked images. This is because some structures
may not be real, but artificially created by the mask-
ing. For example, while Figure 1 (left) superficially sug-
gests that a gap exists in the disc, similar structures in
the masked images of the external companion and self-
gravitating disc in Figure 3 are generated by the inter-
action of the mask with a smoothly declining emissivity
Figure 4. Radial emission profiles for the successful gravitation-
ally unstable disc simulation (red), which has been decreased by a
factor of ∼ 1.5 (see text) and the successful companion disc simu-
lation (blue). Also shown are data from P´erez et al. (2016) (black
points) and the best fit model from Andrews et al. (2009) (black
dash). The steep decline beyond ≈ 300 au is key to matching the
observed morphology in the simulated images.
profile. Masked images of observations therefore need to
be interpreted via the type of forward modelling exercise
undertaken here.
A two-armed spiral structure, as observed around Elias
2-27, is consistent with high disc-to-star mass ratios (e.g.
Lodato & Rice 2005). Indeed Tomida et al. (2017), who
suggested that the Elias 2-27 disc is self-gravitating, ob-
tained a high disc-to-star-mass ratio in their simulations.
The shocks associated with spiral arms in self-gravitating
discs have been shown to have an effect on the chemistry
of the disc material (Ilee et al. 2011; Hincelin et al. 2013;
Evans et al. 2015) leading to the prospect of detecting
the features in line emission (Douglas et al. 2013). Fu-
ture observations of the Elias 2-27 system, with sensitiv-
ities high enough to spatially resolve relevant molecular
line transitions (e.g. CO, HCO+, OCS and H2CO) will
be crucial in further evaluating the dynamics occurring
within the star-disc system.
Finally, deep near infrared imaging should offer a first
step towards deciding which scenario is at work in the
system. If no companion is detected, gravitationally in-
stability is likely. Otherwise further follow-up observa-
tions would be required to confirm any possible detected
companion.
5. CONCLUSIONS
We present the results of a series of hydrodynamic and
radiative transfer models to test three hypotheses regard-
ing the origin of the disc morphology around Elias 2-27
-- a companion internal or external to the spirals, and
a gravitationally unstable disc. Our results show that a
steep decline in surface mass density beyond ≈ 300 au is
required, and that the gravitational instability hypothe-
sis or a (cid:46) 10−13MJup companion between ≈ 300−700 au
can reproduce all components of the observed morphol-
101102Radius [au]10-410-310-210-1100Sº [Jy arcsec-2]G.I. (rescaled)ExternalBest-fit model from Andrews et al. 2009Radial profile of emission from Perez et al. 20166
Meru et al.
Figure 5. Left: surface density of an external companion simulation (see Section 4 for details) which displays prominent two-armed
spirals. This simulation involved a 0.5M(cid:12) companion on a circular orbit at R = 1200 au and was run for 1.3 orbits at 1200au. Right:
unsharp masked image of this simulation, which is unable to reproduce the morphology due to higher levels of emission in the inner disc.
ogy. Given this, we suggest that Elias 2-27 may be
one of the first examples of an observed self-gravitating
disc or a disc that has recently fragmented forming a
(cid:46) 10 − 13MJup planet.
We thank Mike Irwin for obtaining UKIDSS pho-
tometry, and Jim Pringle and the referee for use-
ful comments. We acknowledge support from the
DISCSIM project, grant agreement 341137 under ERC-
2013-ADG. FM acknowledges support from The Lev-
erhulme Trust. This paper uses the following ALMA
data: ADS/JAO.ALMA# 2013.1.00498.S. This work
used the Darwin DiRAC HPC cluster at the Univer-
sity of Cambridge, and the Cambridge COSMOS SMP
system funded by ST/J005673/1, ST/H008586/1 and
ST/K00333X/1 grants.
REFERENCES
Allard, F., Hauschildt, P. H., Alexander, D. R., & Starrfield, S.
1997, ARA&A, 35, 137
ALMA Partnership, Brogan, C. L., P´erez, L. M., et al. 2015, ApJ,
808, L3
Andrews, S. M., Wilner, D. J., Hughes, A. M., Qi, C., &
Dullemond, C. P. 2009, ApJ, 700, 1502
Andrews, S. M., Wilner, D. J., Zhu, Z., et al. 2016, ApJ, 820, L40
Baraffe, I., Chabrier, G., Allard, F., & Hauschildt, P. H. 1998,
A&A, 337, 403
Bate, M. R., Bonnell, I. A., & Price, N. M. 1995, MNRAS, 277,
362
Benz, W. 1990, in Numerical Modelling of Nonlinear Stellar
Pulsations Problems and Prospects,, ed. J. R. Buchler (Kluwer,
Dordrecht), 269
Birnstiel, T., Dullemond, C. P., & Pinilla, P. 2013, A&A, 550, L8
Booth, R. A., & Clarke, C. J. 2016, MNRAS, 458, 2676
Casassus, S., van der Plas, G., M, S. P., et al. 2013, Nature, 493,
191
Clarke, C. J., & Lodato, G. 2009, MNRAS, 398, L6
Crida, A., Morbidelli, A., & Masset, F. 2006, Icarus, 181, 587
Douglas, T. A., Caselli, P., Ilee, J. D., et al. 2013, MNRAS, 433,
2064
Evans, M. G., Ilee, J. D., Boley, A. C., et al. 2015, MNRAS, 453,
1147
Evans, II, N. J., Dunham, M. M., Jørgensen, J. K., et al. 2009,
ApJS, 181, 321
Hauschildt, P. H., Allard, F., & Baron, E. 1999, ApJ, 512, 377
Hincelin, U., Wakelam, V., Commer¸con, B., Hersant, F., &
Guilloteau, S. 2013, ApJ, 775, 44
Ilee, J. D., Boley, A. C., Caselli, P., et al. 2011, MNRAS, 417,
2950
Isella, A., Carpenter, J. M., & Sargent, A. I. 2009, ApJ, 701, 260
Isella, A., Guidi, G., Testi, L., et al. 2016, Phys. Rev. Lett., 117,
251101.
http://link.aps.org/doi/10.1103/PhysRevLett.117.251101
Lawrence, A., Warren, S. J., Almaini, O., et al. 2007, MNRAS,
379, 1599
Lin, D. N. C., & Papaloizou, J. C. B. 1993, in Protostars and
Planets III, ed. E. H. Levy & J. I. Lunine, 749 -- 835
Lodato, G., & Rice, W. K. M. 2005, MNRAS, 358, 1489
Luhman, K. L., & Rieke, G. H. 1999, ApJ, 525, 440
Malik, M., Meru, F., Mayer, L., & Meyer, M. 2015, ApJ, 802, 56
McMullin, J. P., Waters, B., Schiebel, D., Young, W., & Golap,
K. 2007, in Astronomical Society of the Pacific Conference
Series, Vol. 376, Astronomical Data Analysis Software and
Systems XVI, ed. R. A. Shaw, F. Hill, & D. J. Bell, 127
Meru, F. 2015, MNRAS, 454, 2529
Monaghan, J. J. 1992, ARA&A, 30, 543
Morris, J. P., & Monaghan, J. J. 1997, J. Comp. Phys., 136, 41
Natta, A., Testi, L., & Randich, S. 2006, A&A, 452, 245
Paardekooper, S.-J., & Mellema, G. 2006, A&A, 453, 1129
P´erez, L. M., Carpenter, J. M., Andrews, S. M., et al. 2016,
Science, 353, 1519
Price, D. J., & Bate, M. R. 2007, MNRAS, 377, 77
Ratzka, T., Kohler, R., & Leinert, C. 2005, A&A, 437, 611
Ricci, L., Testi, L., Natta, A., & Brooks, K. J. 2010, A&A, 521,
A66
Shi, J.-M., & Chiang, E. 2014, ApJ, 789, 34
Tobin, J. J., Kratter, K. M., Persson, M. V., et al. 2016, Nature,
538, 483
Tomida, K., Machida, M. N., Hosokawa, T., Sakurai, Y., & Lin,
C. H. 2017, ApJ, 835, L11
Weingartner, J. C., & Draine, B. T. 2001, ApJ, 548, 296
Whitehouse, S. C., & Bate, M. R. 2006, MNRAS, 367, 32
Whitehouse, S. C., Bate, M. R., & Monaghan, J. J. 2005,
MNRAS, 364, 1367
Zhu, Z., Nelson, R. P., Dong, R., Espaillat, C., & Hartmann, L.
2012, ApJ, 755, 6
-1-0.500.5log column density [g/cm2]0400-400-4000400200-200-200200x [au]y [au]log Surface Density [g cm-2]0.50-0.5-1EXTERNAL16h26m45.1s45.0s44.9s-24°23'06.0"07.0"08.0"09.0"10.0"R.A.Dec.10-310-210-1Sº (arbitrary units) |
0912.2996 | 1 | 0912 | 2009-12-15T21:00:10 | A single sub-km Kuiper Belt object from a stellar Occultation in archival data | [
"astro-ph.EP",
"astro-ph.SR"
] | The Kuiper belt is a remnant of the primordial Solar System. Measurements of its size distribution constrain its accretion and collisional history, and the importance of material strength of Kuiper belt objects (KBOs). Small, sub-km sized, KBOs elude direct detection, but the signature of their occultations of background stars should be detectable. Observations at both optical and X-ray wavelengths claim to have detected such occultations, but their implied KBO abundances are inconsistent with each other and far exceed theoretical expectations. Here, we report an analysis of archival data that reveals an occultation by a body with a 500 m radius at a distance of 45 AU. The probability of this event to occur due to random statistical fluctuations within our data set is about 2%. Our survey yields a surface density of KBOs with radii larger than 250 m of 2.1^{+4.8}_{-1.7} x 10^7 deg^{-2}, ruling out inferred surface densities from previous claimed detections by more than 5 sigma. The fact that we detected only one event, firmly shows a deficit of sub-km sized KBOs compared to a population extrapolated from objects with r>50 km. This implies that sub-km sized KBOs are undergoing collisional erosion, just like debris disks observed around other stars. | astro-ph.EP | astro-ph |
A single sub-km Kuiper Belt object from a stellar
Occultation in archival data
H. E. Schlichting1,2, E. O. Ofek1,3, M. Wenz4, R. Sari1,5,
A. Gal-Yam6, M. Livio7, E. Nelan7, S. Zucker8
1 Department of Astronomy, 249-17, California Institute of Technology, Pasadena, CA 91125, USA
2 CITA, University of Toronto, 60 St. George St., ON, M5S 3H8, Canada
3 Einstein Fellow
4 Goddard Space Flight Center, 8800 Greenbelt Road, Greenbelt, MD 20771, USA
5 Racah Institute of Physics, Hebrew University, Jerusalem 91904, Israel
6 Faculty of Physics, Weizmann Institute of Science, POB 26, Rehovot 76100, Israel
7 Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218, USA
8 Department of Geophysics and Planetary Sciences, Tel Aviv University, Tel Aviv 69978, Israel
The Kuiper belt is a remnant of the primordial Solar System. Measurements of
its size distribution constrain its accretion and collisional history, and the importance
of material strength of Kuiper belt objects (KBOs)(1; 2; 3; 4). Small, sub-km sized,
KBOs elude direct detection, but the signature of their occultations of background
stars should be detectable(5; 6; 7; 8; 9). Observations at both optical(10) and X-
ray(11) wavelengths claim to have detected such occultations, but their implied KBO
abundances are inconsistent with each other and far exceed theoretical expectations.
Here, we report an analysis of archival data that reveals an occultation by a body with
a ∼500 m radius at a distance of 45 AU. The probability of this event to occur due to
random statistical fluctuations within our data set is about 2%. Our survey yields a
surface density of KBOs with radii larger than 250 m of 2.1+4.8
−1.7 × 107 deg−2, ruling out
inferred surface densities from previous claimed detections by more than 5 σ. The
fact that we detected only one event, firmly shows a deficit of sub-km sized KBOs
compared to a population extrapolated from objects with r > 50 km. This implies that
sub-km-sized KBOs are undergoing collisional erosion, just like debris disks observed
around other stars.
A small KBO crossing the line of sight to a star will partially obscure the stellar light,
an event which can be detected in the star's light curve. For visible light, the characteristic
scale of diffraction effects, known as the Fresnel scale, is given by (λa/2)1/2 ∼ 1.3 km, where
a ∼ 40 AU is the distance to the Kuiper belt and λ ∼ 600 nm is the wavelength of our
observations.
Diffraction effects will be apparent in the star's light curve due to occulting KBOs
provided that both star and the occulting object are smaller than the Fresnel scale (12; 13).
Occultations by objects smaller than the Fresnel scale are in the Fraunhofer regime.
In
this regime the diffraction pattern is determined by the size of the KBO and its distance
-- 2 --
to the observer, the angular size of the star, the wavelength range of the observations and
the impact parameter between the star and the KBO (see Supplementary Information for
details). The duration of the occultation is approximately given by the ratio of the Fresnel
scale to the relative velocity perpendicular to the line of sight between the observer and the
KBO. Since the relative velocity is usually dominated by the Earth's velocity around the
Sun, which is 30 km s−1, typical occultations only last of order of a tenth of a second.
Extensive ground based efforts have been conducted to look for optical occultations
(10; 9; 14; 15). To date, these visible searches have announced no detections in the region
of the Kuiper belt (30-60 AU), but one of these quests claims to have detected some events
beyond 100 AU and at about 15 AU (10). Unfortunately, ground based surveys may suffer
from a high rate of false-positives due to atmospheric scintillation, and lack the stability
of space based platforms. The ground breaking idea to search for occultations in archival
RXTE X-ray data resulted in several claimed occultation events (11). Later, revised analysis
of the X-ray data (16; 17; 18; 19) conclude that the majority of the originally reported events
are most likely due to instrumental dead time effects. Thus, previous reports of optical and
X-ray events remain dubious (14) and their inferred KBO abundance is inconsistent with the
observed break in the KBO size distribution, which has been obtained from direct detections
of large KBOs (20; 21; 22). Furthermore, they are also difficult to reconcile with theoretical
expectations, which predict collisional evolution for KBOs smaller than a few km in size
(23; 4) and hence a lower KBO abundance than inferred from extrapolation from KBOs
with r > 50 km.
For the past 14 years, the Fine Guidance Sensors (FGS ) on board of Hubble Space
Telescope (HST ) have been collecting photometric measurements of stars with 40 Hz time
resolution, allowing for the detection of the occultation diffraction pattern rather than a
simple decrease in the photon count. We examined four and a half years of archival FGS
data, which contain ∼ 12, 000 star hours of low ecliptic latitude (b < 20◦) observations.
Our survey is most likely to detect occultations by KBOs that are 200-500 m in radius
given the signal-to-noise of our data (Supplementary Figure 3) and a power-law size distri-
bution with power-law index between 3 and 4.5. Occultation events in this size range are
in the Fraunhofer regime where the depth of the diffraction pattern varies linearly with the
area of the occulting object and is independent of its shape. The theoretical light curves for
our search algorithm were therefore calculated in this regime. We fitted these theoretical
occultation templates to the FGS data and performed χ2 analysis to identify occultation
candidates (see Supplementary Information). We detected one occultation candidate, at
ecliptic latitude 14◦, that significantly exceeds our detection criterion (Figure 1). The best
fit parameters yield a KBO size of r = 520 ± 60 m and a distance of 45+5
−4 AU where we
assumed a circular KBO orbit and an inclination of 14◦. Using bootstrap simulations, we
estimate a probability of ∼ 2% that such an event is caused by statistical fluctuations over
the whole analyzed FGS data set (Supplementary Figure 7). We note that for objects on
circular orbits around the sun two solutions can fit the duration of the event. However, the
-- 3 --
other solution is at a distance of 0.07 AU from the Earth, and is therefore unlikely. It is
also unlikely that the occulting object was located in the Asteroid belt, since the expected
occultation rate from Asteroids is about two orders of magnitude less than our implied rate.
Furthermore, an Asteroid would have to have an eccentricity of order unity to be able to
explain the duration of the observed occultation event.
Using the KBO ecliptic latitude distribution from Elliot et. al (2005) (24), our detection
efficency, and our single detection, we constrain the surface density around the ecliptic
(averaged over −5◦ < b < 5◦) of KBOs with radii larger than 250 m to 2.1+4.8
−1.7 × 107 deg−2
(see Supplementary Information Sections 5 and 6). This surface density is about three
times the implied surface density at 5.5◦ ecliptic latitude and about five times the surface
density at 8 − 20◦ ecliptic latitude. This is the first measurement of the surface density of
hecto-meter-sized KBOs and it improves previous upper limits by more than an order of
magnitude (9; 15). Figure 2 displays our measurement for the sub-km KBO surface density
and summarizes published upper limits from various surveys. Our original data analysis
focused on the detection of KBOs located at the distance of the Kuiper belt between 30 AU
and 60 AU. In order to compare our results with previously reported ground-based detections
beyond 100 AU (10), we performed a second search of the FGS data that was sensitive to
objects located beyond the classical Kuiper belt. Our results challenge the reported ground-
based detections of two 300 m-sized objects beyond 100 AU (10). Given our total number of
star hours and a detection efficency of 3% for 300 m-sized objects at ∼ 100 AU we should have
detected more than twenty occultations. We therefore rule out the previously claimed optical
detections (10) by more than 5 σ. This result accounts for the broad latitude distribution of
our observations (i.e., b < 20◦) and the quoted detection efficency of our survey includes
the effect of the finite angular radii of the guide stars at 100 AU.
The KBO cumulative size distribution is parameterized by N(> r) ∝ r1−q, where N(>
r) is the number of objects with radii greater than r, and q is the power-law index. The
power-law index for KBOs with radii above ∼45 km is ∼ 4.5 (21; 22) and there is evidence
for a break in the size distribution at about rbreak ∼ 45 km (20; 21; 22). We hence use this
break radius and assume a surface density for KBOs larger than rbreak (25) of 5.4 deg−2
around the ecliptic. Accounting for our detection efficency, the velocity distribution of the
HST observations, and assuming a single power-law for objects with radii less than 45 km
in size, we find q = 3.9+0.3,+0.4
−0.3,−0.7 (1 and 2 σ errors) below the break. Our results firmly show a
deficit of sub-km sized KBOs compared to large objects. This confirms the existence of the
previously reported break and establishes a shallower size distribution extending two orders
of magnitude in size down to sub-km sized objects. This suggests that sub-km sized KBOs
underwent collisional evolution, eroding the smaller KBOs. This collisional grinding in the
Kuiper belt provides the missing link between large KBOs and dust producing debris disks
around other stars. Currently our results are consistent with a power-law index of strength
dominated collisional cascade (23), q = 3.5, within 1.3σ and with predictions for strengthless
rubble piles (4), q = 3.0, within 2.4σ. An intermediate value of 3 < q < 3.5 implies that
-- 4 --
KBOs are strengthless rubble piles above some critical size, rc < r < 45 km, and strength
dominated below it, r < rc. Our observations constrain for the first time rc. At the 2 σ level
we find rc > 3 km.
Using our estimate for the size distribution power-law index (q = 3.9) and our KBO
surface density for 250 m sized KBOs at an ecliptic latitude of b = 5.5◦, which is the ecliptic
latitude of the RXTE observations of Scorpius X-1, we predict that there should be ∼
3.6 × 109 30 m-radius objects per square degree. This is about 150 times less than the
original estimate from X-ray observations of Scorpius X-1 that reported 58 events (11), and
it is about 30 times less than the revised estimate from the same X-ray observations, which
concludes that up to 12 events might be actual KBO occultations (16). Our results rule out
the implied surface density from these 12 events at 7 σ confidence level. One can reconcile
our results and the claimed X-ray detections only by invoking a power-law index of q ∼ 5.5
between 250 m and 30 m. More recent X-ray work reports no new detections in the region of
the Kuiper belt but places an upper limit of 1.7 ×1011 deg−2 for objects of 50 m in radius and
larger (18). This is consistent with the KBO surface density of N(> 50 m) = 8.2 × 108 deg−2
that we derive by extrapolating from our detection in the hecto-meter size range.
The statistical confidence level on our detection is 98%. However, our conclusions that
there is a significant break in the size distribution and that collisional erosion is taking place
and the significant discrepancy with previously claimed occultation detections rely on the low
number of events we discovered. These conclusions would only be strengthened if this event
was caused by an unlikely statistical fluctuation or a yet unknown instrumental artifact.
Ongoing analysis of the remaining FGS data, which will triple the number of star hours,
together with further development of our detection algorithm (i.e., including a larger number
of light curve templates) holds the promise for additional detections of occultation events
and will allow us to constrain the power-law index of the size distribution further.
-- 5 --
REFERENCES
(1) Davis, D. R. & Farinella, P. Collisional Evolution of Edgeworth-Kuiper Belt Objects.
Icarus 125, 50 -- 60 January 1997.
(2) Stern, S. A. & Colwell, J. E. Collisional Erosion in the Primordial Edgeworth-Kuiper
Belt and the Generation of the 30-50 AU Kuiper Gap. ApJ 490, 879 -- 882 December
1997.
(3) Kenyon, S. J. & Luu, J. X. Accretion in the Early Kuiper Belt. II. Fragmentation. AJ
118, 1101 -- 1119 August 1999.
(4) Pan, M. & Sari, R. Shaping the Kuiper belt size distribution by shattering large but
strengthless bodies. Icarus 173, 342 -- 348 February 2005.
(5) Bailey, M. E. Can 'invisible' bodies be observed in the solar system. Nature 259, 290 -- 291
January 1976.
(6) Dyson, F. J. Hunting for comets and planets. QJRAS 33, 45 -- 57 June 1992.
(7) Axelrod, T. S., Alcock, C., Cook, K. H. & Park, H.-S. in Robotic Telescopes in the 1990s
(ed Filippenko, A. V.) 171 -- 181 1992).
(8) Roques, F., Moncuquet, M. & Sicardy, B. Stellar occultations by small bodies - Diffrac-
tion effects. AJ 93, 1549 -- 1558 June 1987.
(9) Zhang, Z.-W., Bianco, F. B., Lehner, M. J., Coehlo, N. K., Wang, J.-H. et al. First
Results from the Taiwanese-American Occultation Survey (TAOS). ApJ 685, L157 --
L160 October 2008.
(10) Roques, F., Doressoundiram, A., Dhillon, V., Marsh, T., Bickerton, S. et al. Exploration
of the Kuiper Belt by High-Precision Photometric Stellar Occultations: First Results.
AJ 132, 819 -- 822 August 2006.
(11) Chang, H.-K., King, S.-K., Liang, J.-S., Wu, P.-S., Lin, L. C.-C. et al. Occultation of
X-rays from Scorpius X-1 by small trans-neptunian objects. Nature 442, 660 -- 663
August 2006.
(12) Roques, F. & Moncuquet, M. A Detection Method for Small Kuiper Belt Objects: The
Search for Stellar Occultations. Icarus 147, 530 -- 544 October 2000.
(13) Nihei, T. C., Lehner, M. J., Bianco, F. B., King, S.-K., Giammarco, J. M. et al.
Detectability of Occultations of Stars by Objects in the Kuiper Belt and Oort Cloud.
AJ 134, 1596 -- 1612 October 2007.
-- 6 --
(14) Bickerton, S. J., Kavelaars, J. J. & Welch, D. L. A Search for SUB-km Kuiper Belt
Objects with the Method of Serendipitous Stellar Occultations. AJ 135, 1039 -- 1049
March 2008.
(15) Bianco, F. B., Protopapas, P., McLeod, B. A., Alcock, C. R., Holman, M. J. et al. A
Search for Occultations of Bright Stars by Small Kuiper Belt Objects using Megacam
on the MMT. AJ 138, 568 -- 578 August 2009.
(16) Chang, H.-K., Liang, J.-S., Liu, C.-Y. & King, S.-K. Millisecond dips in the RXTE/PCA
light curve of Sco X-1 and trans-Neptunian object occultation. MNRAS 378, 1287 --
1297 July 2007.
(17) Jones, T. A., Levine, A. M., Morgan, E. H. & Rappaport, S. Production of Millisecond
Dips in Sco X-1 Count Rates by Dead Time Effects. ApJ 677, 1241 -- 1247 April 2008.
(18) Liu, C.-Y., Chang, H.-K., Liang, J.-S. & King, S.-K. Millisecond dip events in the
2007 RXTE/PCA data of Sco X-1 and the trans-Neptunian object size distribution.
MNRAS 388, L44 -- L48 July 2008.
(19) Blocker, A. W., Protopapas, P.& Alcock, C. R. A Bayesian Approach to the Analysis
of Time Symmetry in Light Curves: Reconsidering Scorpius X-1 Occultations. ApJ
701, 1742 -- 1752 August 2009.
(20) Bernstein, G. M., Trilling, D. E., Allen, R. L., Brown, M. E., Holman, M. et al. The
Size Distribution of Trans-Neptunian Bodies. AJ 128, 1364 -- 1390 September 2004.
(21) Fuentes, C. I. & Holman, M. J. a SUBARU Archival Search for Faint Trans-Neptunian
Objects. AJ 136, 83 -- 97 July 2008.
(22) Fraser, W. C., Kavelaars, J. J., Holman, M. J., Pritchet, C. J., Gladman, B. J. et al.
The Kuiper belt luminosity function from m(R)=21 to 26. Icarus 195, 827 -- 843 June
2008.
(23) Dohnanyi, J. W. Collisional models of asteroids and their debris. J. Geophys. Res. 74,
2531 -- 2554 (1969).
(24) Elliot, J. L., Kern, S. D., Clancy, K. B., Gulbis, A. A. S., Millis, R. L. et al. The
Deep Ecliptic Survey: A Search for Kuiper Belt Objects and Centaurs. II. Dynamical
Classification, the Kuiper Belt Plane, and the Core Population. AJ 129, 1117 -- 1162
February 2005.
(25) Fuentes, C. I., George, M. R. & Holman, M. J. A Subaru Pencil-Beam Search for m(R)
∼ 27 Trans-Neptunian Bodies. ApJ 696, 91 -- 95 May 2009.
(26) Skrutskie, M. F., Cutri, R. M., Stiening, R., Weinberg, M. D., Schneider, S. et al. The
Two Micron All Sky Survey (2MASS). AJ 131, 1163 -- 1183 February 2006.
-- 7 --
(27) Monet, D. G., Levine, S. E., Canzian, B., Ables, H. D., Bird, A. R. et al. The USNO-B
Catalog. AJ 125, 984 -- 993 February 2003.
(28) Roques, F., Georgevits, G. & Doressoundiram, A. The Kuiper Belt Explored by
Serendipitous Stellar Occultations. The University of Arizona Press, 545 -- 556 2008.
Acknowledgments We thank Dr. H. K. Chang for valuable comments that helped to im-
prove this manuscript. Some of the numerical calculations presented here were performed on
Caltech's Division of Geological and Planetary Sciences Dell cluster. Partial support for this
research was provided by NASA through a grant from the Space Telescope Science Institute.
R. S. acknowledges support from the ERC and the Packard Foundation. A. G. is supported
by the Israeli Science Foundation, an EU Seventh Framework Programme Marie Curie IRG
fellowship and the Benoziyo Center for Astrophysics, a research grant from the Peter and
Patricia Gruber Awards, and the William Z. and Eda Bess Novick New Scientists Fund at
the Weizmann Institute. S. Z. acknowledges support from the Israel Science Foundation --
Adler Foundation for Space Research.
Author Contributions H. E. S. wrote the detection algorithm, analyzed the FGS data for
occultation events, calculated the detection efficency of the survey, preformed the bootstrap
analysis and wrote the paper. E. O. O. calculated the stellar angular radii, the velocity
information of the observations, the correlated noise and other statistical properties of the
data. R. S. guided this work and helped with the scientific interpretation of the results. A.G.
proposed using HST FGS data for occultation studies and helped to make the data avail-
able for analysis. M. W. extracted the FGS photometry streams and provided coordinates
and magnitudes of the guide stars. M. L. helped in gaining access to the FGS data and
provided insights into the operation and noise properties of the FGS . E. N. provided expert
interpretation of the FGS photometric characteristics in the HST operational environment.
S. Z. took part in the statistical analysis of the data. All authors discussed the results and
commented on the manuscript.
Author Information Reprints and permissions information is available at www.nature.com/reprints.
Correspondence and requests for materials should be addressed to H. E. S. ([email protected])
or E. O. O. ([email protected]).
This preprint was prepared with the AAS LATEX macros v5.2.
-- 8 --
a
1.5
2
b
-0.5
-1
1
time relative to mid eclipse (s)
0
0.5
-2
-1.5
t
n
u
o
c
t
n
o
o
h
p
t
n
u
o
c
n
o
t
o
h
p
220
200
180
160
140
120
100
80
220
200
180
160
140
120
100
80
-0.3
-0.2
-0.1
0
0.1
0.2
0.3
time relative to mid eclipse (s)
Fig. 1. -- Photon counts as a function of time of the candidate occultation event observed by FGS2. Part
a) shows the photon count spanning ±2 seconds around the occultation event. Part b) displays the event
in detail. The red crosses and error bars are the FGS data points with Poisson error bars, the dashed,
blue line is the theoretical diffraction pattern (calculated for the 400-700 nm wavelength range of the FGS
observations), and the pink squares correspond to the theoretical light curve integrated over 40 Hz intervals.
Note, the actual noise for this observation is about 4% larger than Poisson noise due to additional noise
sources such as dark counts (about 3 to 6 counts in a 40 Hz interval), and jitter due to the displacement of
the guide star (by up to 10 mas) from its null position. The mean signal-to-noise ratio in a 40 Hz interval
for the roughly half an hour of observations is ∼ 12. The event occurred at UTC 05:17:49 2007, Mar 24.
The best fit χ2/dof is 20.1/21. The star has an ecliptic latitude of +14. Its angular radius and effective
temperature are ≈ 0.3 of the Fresnel scale and ≈ 4460 K, respectively. These values were derived by fitting
the 2MASS (26) JHK and USNO-B1 BR (27) photometry with a black-body spectrum. The position of
the star is R.A.=186.87872◦, Dec=12.72469◦ (J2000) and its estimated V-magnitude is 13.4. The auto-
correlation function (excluding lag zero) of the photometric time series of this event is consistent with zero
within the statistical uncertainty. Each FGS provides two independent PMT readings and we confirmed that
the occultation signature is present in both of these independent photon counts. We examined the photon
counts of the other guide star that was observed by FGS1 at the time of the occultation and confirmed that
the occultation signal is only present in the observations recorded by FGS2. We examined the engineering
telemetry for HST around the time of the event and verified that the guiding performance of HST was
normal. We therefore conclude that the above occultation pattern is not caused by any known instrumental
artifacts.
-- 9 --
FGS
FGS
RXTE Jones et al. (2008)
RXTE Liu et al. (2008)
RXTE Chang et al. (2007)
TAOS Zhang et al. (2008)
MMT Bianco et al. (2009)
Bickerton et al. (2008)
Roques et al. (2008)
12
12
10
10
8
8
6
6
4
4
2
2
)
)
2
2
-
-
g
g
e
e
d
d
(
(
)
)
r
r
>
>
(
(
N
N
g
g
o
o
l
l
0
0
0.01
0.01
0.1
0.1
1
1
radius (km)
radius (km)
10
10
100
100
Fig. 2. -- Cumulative KBO size distribution as a function of KBO radius for objects located between 30 and
60 AU. The results from our FGS survey are shown in red and are presented in three different ways: (i) The
cross is derived from our detection and represents the KBO surface density around the ecliptic (averaged over
−5◦ < b < 5◦) and is shown with 1 σ error bars. The cross is plotted at r = 250 m, which is roughly the peak
of our detection probability (see Supplementary Information Section 6 for details). (ii) The upper and lower
red curves correspond to our upper and lower 95% confidence level which were derived without assuming any
size distribution. (iii) The region bounded by the two straight red lines falls within 1σ of our best estimate
for the power-law size distribution index, i.e. q = 3.9 ± 0.3, which was calculated for low ecliptic latitudes
(b < 5◦). These lines are anchored to the observed surface density at r = 45 km. For comparison, the
green (long-dashed) line is the observed size distribution of large KBOs (i.e., r > 45 km), which has q = 4.5,
extrapolated as a single power-law to small sizes. The blue (short-dashed) line is a double power-law with
q = 3.5 (collisional cascade of strength dominated bodies) for KBOs with radii less than 45 km and q = 4.5
above. The cyan (dot-dashed) line corresponds to q = 3.0 (collisional cascade of strengthless rubble piles)
for KBOs below 45 km in size. All distributions are normalized to N (> r) = 5.4 deg−2 at a radius of 45 km
(25). In addition, 95% upper limits from various surveys are shown in black. Note, a power-law index of 3.9
was used for calculating the cumulative KBO number density from the RXTE observations.
-- 10 --
Supplementary Information
1 The FGS data set
There are three FGS on board of HST . Each FGS consists of four photomultipliers (PMTs).
Nominal HST operation uses two FGS for guiding, with each FGS observing its own guide
star. The photon counts recorded by each FGS are therefore different, but global instru-
mental artifacts and Observatory level transients will display in both FGS and can therefore
be identified and removed.
Observations of the inclination distribution of large KBOs find that about 75% have an
inclination angle i . 20◦ (29; 30; 31). We therefore divide the FGS observations into a low
ecliptic latitude (b < 20◦) and a high ecliptic latitude (b > 20◦) sample. The high-ecliptic
latitude observations (b > 20◦) provide an excellent control sample.
2 FGS Guide Stars
The FGS guide stars span a broad range of magnitudes and spectral types. The signal-to-
noise ratio, S/N, in a 1/40 s data bin depends on the magnitude of the star. Its distribution
is shown in Supplementary Figure 3.
The angular sizes of guide stars were derived by fitting the 2MASS (32) JHK and USNO-
B1 BR (33) photometry with a black-body spectrum. Supplementary Figure 4 shows the
angular radii distribution of the guide stars. About 66% of the stars in our data set subtend
angular sizes less than 0.5 of the Fresnel scale at a distance of 40 AU. The diffraction pattern
that is produced by a sub-km sized KBO occulting an extended background star is smoothed
over the finite stellar disk. This effect becomes clearly noticeable for stars that subtend sizes
larger than about 0.5 of a Fresnel scales (34; 35) and it reduces the detectability of occul-
tation events around such stars. The effect of finite angular radii of the guide stars on the
detection efficiency of our survey is taken into account (see Detection Efficency section 5 for
details).
3 Detection Algorithm
Our detection algorithm performs a template search with theoretical light curves and uses a
χ2 fitting procedure to identify occultation candidates. Our survey is most likely to detect
KBO occultation events caused by objects that are 200-500 m in radius given the signal-
to-noise of our data and for a power-law index of the KBO size distribution, q, between 3
and 4.5. Occultation events in this size range are in the Fraunhofer regime. The theoretical
-- 11 --
light curves for our search algorithm are therefore calculated in the Fraunhofer regime. Our
templates are calculated for various impact parameters assuming a point source background
star and are integrated over the 400-700 nm wavelength range of the FGS observations. For
a given impact parameter between the KBO and the star, our theoretical light curves have
three free parameters that we fit for. The first is the mean number of photon counts, which
is the normalization of the light curve. The second is the amplitude of the occultation, which
is proportional to the size of the KBO, and the third is the width of the occultation, which
is independent of the object size, and is determined by the ratio of the Fresnel scale to the
relative speed between HST and the KBO perpendicular to the line of sight. This relative
speed is determined by the combination of HST 's velocity around the Earth, Earth's veloc-
ity around the Sun and the velocity of the KBO itself. We use this information to restrict the
parameter space for the template widths in our search such that we are sensitive to KBOs
located at the distance of the Kuiper belt between 30 AU and 60 AU.
4 Detection Criterion and Significance Estimates
The significance of occultation candidates can be measured by their ∆χ2 which is defined
here as the difference between the χ2 calculated for the best fit of a flat light curve, which
corresponds to no event, and the χ2 of the best fit template. Occultation events have large
∆χ2, since they are poorly fit by a constant. Cosmic ray events, which give rise to one very
large photon count reading in a 40 Hz interval, can also result in a large ∆χ2 but the fit of
the occultation template is very poor. We examined all flagged events for which the tem-
plate fit of the diffraction pattern was better than 15 σ. About a handful of false-positives
where flagged by our detection algorithm that have a value of ∆χ2 comparable to or larger
than the occultation event. However, in all cases these false-positives were caused by a 1 Hz
jitter due to the displacement of the guide star from its null position. The occultation event
itself did not show any such jitter. To determine the ∆χ2 detection criterion for our search
algorithm and to estimate the probability that detected events are due to random noise we
use the bootstrap technique (36). Specifically, from a given FGS time series of length N we
randomly selected N points with repetitions and created 'artificial' time series from it. We
analyzed these 'artificial' data sets using the same search algorithm that we applied to the
actual FGS data. This technique creates random time series with noise properties identical
to those of the actual data, but it will lose any correlated noise. Therefore, this technique
is justified if there is no correlated noise in the data sets. To look for correlated noise we
calculated the auto-correlation function, with lags between 0 to 1 s. Most of the data sets
are free of statistical significant correlated noise. The ∼ 12% of the data sets that did show
correlated noise exceeding 4 σ, which was often due to slopes (e.g., long-term variability) in
the data sets, were excluded from the bootstrap analysis.
The FGS data set consists of observations of many different stars with magnitudes rang-
-- 12 --
ing from 9 to 14. The number of photon counts and signal-to-noise properties vary therefore
from observation to observation (see Supplementary Figure 3 for the signal-to-noise ratio
distribution of the FGS observations). Our ∆χ2 calculation accounts for the Poisson noise
of the data. Therefore, the probability that occultation candidates are due to random noise
can be characterized by a single value of ∆χ2 for all observations, irrespective of the mean
photon count of a given observation provided that the noise properties across all observa-
tions are well characterized by a Poisson distribution. In reality, the noise properties are
different from observation to observation; especially non-Poisson tails in the photon counts
distribution will give rise to slightly different ∆χ2 distributions. Therefore, ideally, we would
determine a unique detection criterion for each individual data set. However, this would re-
quire to simulate each data set, which contains about an hour of observations in a single HST
orbit, over the entire length of our survey (∼ 12, 000 star hours). This is not feasible due to
the enormous computational resources that would be required, i.e. simulating a single one
hour data set over the entire survey length requires about 5 CPU days, which corresponds
to ∼ 60, 000 CPU days for the entire FGS survey. Instead, we perform the bootstrap sim-
ulation over all the FGS data sets together, where each individual data set was simulated
about a 100 times, which required about ∼ 500 CPU days in total. This way we estimate the
typical ∆χ2 value that corresponds to having less than one false-positive detection over the
∼ 12, 000 star hours of low ecliptic observations. For all occultation candidates that exceed
this detection threshold, we determined their statistical significance, i.e. the probability that
they are due to random noise, by extensive bootstrap simulations of the individual data sets
(Supplementary Figure 7).
5 Detection Efficency
The ability to detect an occultation event of a given size KBO depends on the impact
parameter of the KBO, the duration of the event, the angular size of the star and the signal-
to-noise ratio of the data. We determined the detection efficiency of our survey by recovering
synthetic events that we planted into the observed photometric time series by multiplying
the actual FGS data with theoretical light curves of KBO occultation events. The synthetic
events correspond to KBO sizes ranging from 130 m < r < 650 m, they have impact pa-
rameters from 0 to 5.5 Fresnel scales and a relative velocity distribution that is identical to
that of the actual FGS observations. To account for the finite angular sizes of the stars we
generated light curve templates with stellar angular radii of 0.1, 0.2, 0.3, 0.4, 0.6, 0.8 and
1 Fresnel scales distributed as shown in Supplementary Figure 4. The modified light curves
with the synthetic events were analyzed using the same search algorithm that we used to
analyze the FGS data. The detection efficency of our survey was calculated using the angu-
lar size distribution of the FGS guide stars assuming a distance of 40 AU. We normalize our
detection efficency for a given size KBO, η(r) , to 1 for an effective detection cross section
with a radius of one Fresnel scale.
-- 13 --
The detection efficiency of our survey is ∼ 0.05 (∼ 0.6) for objects with r = 200 m
(r = 500 m) located at 40 AU. Note that this value for the detection efficency already ac-
counts for the angular radii distribution of the guide stars (e.g., for comparison, stars that
subtend angular radii less than 0.5 of the Fresnel scale result in a detection efficency of
∼ 0.08 [∼ 0.8] for objects with r = 200 m [r = 500 m].).
6 Calculating the KBO Surface Density
The number of occultation events is given by
Nevents ≃ −2vrelF Z rmax
rmin Z b
−b
η(r)
∆t
∆b
dN(r, b)
dr
dbdr
(1)
where vrel = 23 km/s is the typical relative velocity between the KBO and the observer,
b is the ecliptic latitude, ∆t/∆b is the time observed per degree in ecliptic latitude (see
Supplementary Figure 5) and F = 1.3 km is the Fresnel scale. The number density of KBOs
is both a function of ecliptic latitude and the KBO radius, r. Here we assume that the KBO
latitude distribution, f (b), is independent of size and we take the distribution provided in
Elliot et al. (2005) (31). We further assume that the KBO size distribution follows a power
law. It can therefore be written as N(r, b) = n0 × r−q+1 × f (b) where n0 is the normalization
factor for the cumulative surface density of KBOs. Substituting for dN(r, b)/dr in equation
1 and solving for n0 we have
n0 ≃
Nevents
2vrelF (q − 1)R rmax
rmin
η(r)r−q drR b
−b f(b) ∆t
∆b db
.
(2)
Evaluating equation 2 yields a cumulative KBO surface density averaged over the ecliptic
(b < 5◦) of
N(r > 250 m) ≃ 2.1 × 107 deg−2
(3)
We assumed q = 4 when evaluating the integral over r. We note however that the value for
the cumulative KBO surface density at r = 250 m only depends weakly on the exact choice
for q [e.g. N(r > 250 m) only ranges from 2.3 × 107 deg−2 to 2.1 × 107 deg−2 for values of
q between 3 and 4.5]. We quote our results as the KBO surface density of objects larger
than 250 m in radius since this is roughly the size of KBOs, which our survey is most likely
to detect given our detection efficency and a power-law size distribution with q = 3 − 4.5.
The implied surface density for KBOs with radii larger than 250 m is 7.7 × 106 deg−2 at
b = 5.5◦, which is the ecliptic latitude of the RXTE observations of Scorpius X-1, and it is
4.4 × 106 deg−2 for 8◦ < b < 20◦.
-- 14 --
s
r
u
o
h
r
a
t
s
4000
3500
3000
2500
2000
1500
1000
500
0
0
20
40
60
80
100
S/N
Fig. 3. -- Distribution of star hours as a function of the mean signal-to-noise ratio, S/N , in a 40 Hz bin for
the 12,000 hours of low ecliptic latitude observations (b < 20◦) in the analyzed FGS data set.
-- 15 --
s
r
u
o
h
r
a
t
s
4000
3500
3000
2500
2000
1500
1000
500
0
0
0.5
1
1.5
2
angular radius (Fresnel scale)
Fig. 4. -- Distribution of star hours as a function of angular radii of the guide stars. The angular radii are
given as fraction of the Fresnel scale both which are calculated at 40 AU.
-- 16 --
s
r
u
o
h
r
a
t
s
1200
1000
800
600
400
200
0
-18
-14
-10
-6
-2
2
b (deg)
6
10
14
18
Fig. 5. -- Distribution of star hours as a function of ecliptic latitude, b, for the 12,000 hours of low ecliptic
latitude observations (b < 20◦) in the analyzed FGS data set.
-- 17 --
)
1
-
g
e
d
(
y
t
i
l
i
b
a
b
o
r
p
n
o
i
t
c
e
t
e
d
0.18
0.16
0.14
0.12
0.1
0.08
0.06
0.04
0.02
0
-18
-14
-10
-6
-2
2
b (deg)
6
10
14
18
Fig. 6. -- Detection probability as a function of ecliptic latitude, b, for the 12,000 hours of low ecliptic
latitude observations (b < 20◦) of the analyzed FGS data set. The detection probability was calculated
from the ecliptic latitude distribution of FGS guide stars shown in Supplementary Figure 5 and the KBO
ecliptic latitude distribution from Elliot et al. (2005)(31). Note, we assumed that the KBO ecliptic latitude
distribution is symmetric about the ecliptic and ignored the small ∼ 1.6◦ inclination of the Kuiper belt
plane(31) relative to the ecliptic. For our survey, there is ∼ 60% probability that KBO occultations will occur
inside the low-inclination core region (b < 4◦) of the Kuiper belt. The probability for KBO occultations
outside the core region is roughly uniform for 4◦ < b < 20◦ and about 40% of all KBO occultations will
occur outside the low-inclination core region. The detection of one object at 14◦ is therefore consistent with
the latitude distribution of our observations and that of KBOs.
-- 18 --
10
1
0.1
0.01
)
2
χ
∆
>
(
p
-
f
N
50
54
58
62
∆χ2
66
70
74
Fig. 7. -- Cumulative number of false-positives, Nf −p, as a function of ∆χ2. These false-positives were
obtained from bootstrap simulations using data from ∼ 28 minutes of FGS observations that were acquired
over one HST orbit, in which we discovered the occultation candidate. The original time series was 32
minutes long and we removed the last 4 minutes that showed a significant increasing trend in the number
of photon counts. We removed the occultation event itself (which occurred about 2.3 minutes before the
start of the trend) and simulated 2.5 × 106 star hours, which is 206 times larger than our low ecliptic
latitude observations. This calculation required ∼ 1400 CPU days of computing power. The number of
false-positives, Nf −p, was normalized to 12,000 star hours, which corresponds to the length of the entire
low ecliptic latitude observations. In the entire bootstrap analysis we obtained 4 events with a ∆χ2 ≥ 67.3.
This implies a probability of 8 × 10−7 that events like the occultation candidate with ∆χ2 = 67.3 are caused
by random statistical fluctuations within the original 32 minutes data set that contained the event and a
probability of ∼ 4/206 ∼ 2% that events like the occultation candidate are caused by random statistical
fluctuations over the entire low ecliptic latitude observations. The analysis of our high ecliptic latitude
control sample, which is twice as large, did not yield any events that were comparable in significance to the
occultation candidate.
-- 19 --
REFERENCES
(29) Jewitt, D., Luu, J. & Chen, J. The Mauna Kea-Cerro-Tololo (MKCT) Kuiper Belt and
Centaur Survey. AJ 112, 1225 -- 1238 September 1996.
(30) Brown, M. E. The Inclination Distribution of the Kuiper Belt. AJ 121, 2804 -- 2814 May
2001.
(31) Elliot, J. L., Kern, S. D., Clancy, K. B., Gulbis, A. A. S., Millis, R. L. et al. The
Deep Ecliptic Survey: A Search for Kuiper Belt Objects and Centaurs. II. Dynamical
Classification, the Kuiper Belt Plane, and the Core Population. AJ 129, 1117 -- 1162
February 2005.
(32) Skrutskie, M. F., Cutri, R. M., Stiening, R., Weinberg, M. D., Schneider, S. et al. The
Two Micron All Sky Survey (2MASS). AJ 131, 1163 -- 1183 February 2006.
(33) Monet, D. G., Levine, S. E., Canzian, B., Ables, H. D., Bird, A. R. et al. The USNO-B
Catalog. AJ 125, 984 -- 993 February 2003.
(34) Roques, F. & Moncuquet, M. A Detection Method for Small Kuiper Belt Objects: The
Search for Stellar Occultations. Icarus 147, 530 -- 544 October 2000.
(35) Nihei, T. C., Lehner, M. J., Bianco, F. B., King, S.-K., Giammarco, J. M. et al.
Detectability of Occultations of Stars by Objects in the Kuiper Belt and Oort Cloud.
AJ 134, 1596 -- 1612 October 2007.
(36) Efron, B. The Jackknife, the Bootstrap and other resampling plans. Society for Industrial
Mathematics (1982).
This preprint was prepared with the AAS LATEX macros v5.2.
|
1601.03713 | 1 | 1601 | 2016-01-14T20:11:05 | Surface Albedo and Spectral Variability of Ceres | [
"astro-ph.EP"
] | Previous observations suggested that Ceres has active but possibly sporadic water outgassing, and possibly varying spectral characteristics in a time scale of months. We used all available data of Ceres collected in the past three decades from the ground and the Hubble Space Telescope, and the newly acquired images by Dawn Framing Camera to search for spectral and albedo variability on Ceres, in both a global scale and local regions, particularly the bright spots inside Occator crater, over time scales of a few months to decades. Our analysis has placed an upper limit on the possible temporal albedo variation on Ceres. Sporadic water vapor venting, or any possibly ongoing activity on Ceres, is not significant enough to change the albedo or the area of the bright features in Occator crater by >15%, or the global albedo by >3% over various time scales that we searched. Recently reported spectral slope variations can be explained by changing Sun-Ceres-Earth geometry. The active area on Ceres is less than 1 km$^2$, too small to cause global albedo and spectral variations detectable in our data. Impact ejecta due to impacting projectiles of tens of meters in size like those known to cause observable changes to the surface albedo on Asteroid Scheila cannot cause detectable albedo change on Ceres due to its relatively large size and strong gravity. The water vapor activity on Ceres is independent of Ceres' heliocentric distance, rulling out the possibility of comet-like sublimation process as a possible mechanism driving the activity. | astro-ph.EP | astro-ph | Surface Albedo and Spectral Variability of Ceres
Jian-Yang Li (李荐扬)1, Vishnu Reddy1, Andreas Nathues2, Lucille Le Corre1, Matthew R. M.
Izawa3,4, Edward A. Cloutis3, Mark V. Sykes1, Uri Carsenty5, Julie C. Castillo-Rogez6, Martin
Hoffmann2, Ralf Jaumann5, Katrin Krohn5, Stefano Mottola5, Thomas H. Prettyman1, Michael
Schaefer2, Paul Schenk7, Stefan E. Schröder5, David A. Williams8, David E. Smith9, Maria T.
Zuber10, Alexander S. Konopliv6, Ryan S. Park6, Carol A. Raymond6, Christopher T. Russell11
…
1 Planetary Science Institute, 1700 E. Ft. Lowell Rd., Suite 106, Tucson, AZ 85719, USA
2 Max Planck Institute for Solar System Research, 37077 Göttingen, Germany
3 University of Winnipeg, Winnipeg, Canada
4 Royal Ontario Museum, Toronto, Canada
5 German Aerospace Center (DLR), Institute of Planetary Research, Berlin, Germany
6 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA 91109, USA
7 Lunar and Planetary Institute, Houston, TX 77058, USA
8 School of Earth and Space Exploration, Arizona State University, Tempe, AZ 85287, USA
9 Solar System Exploration Division, NASA Goddard Space Flight Center, Greenbelt, MD
20771, USA
10 Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of
Technology, Cambridge, MA 02139, USA
11 Institute of Geophysics and Planetary Physics, University of California, Los Angeles, CA
90095, USA
1
Abstract
Previous observations suggested that Ceres has active but possibly sporadic water
outgassing, and possibly varying spectral characteristics in a time scale of months. We used all
available data of Ceres collected in the past three decades from the ground and the Hubble Space
Telescope, and the newly acquired images by Dawn Framing Camera to search for spectral and
albedo variability on Ceres, in both a global scale and local regions, particularly the bright spots
inside Occator crater, over time scales of a few months to decades. Our analysis has placed an
upper limit on the possible temporal albedo variation on Ceres. Sporadic water vapor venting, or
any possibly ongoing activity on Ceres, is not significant enough to change the albedo or the area
of the bright features in Occator crater by >15%, or the global albedo by >3% over various time
scales that we searched. Recently reported spectral slope variations can be explained by
changing Sun-Ceres-Earth geometry. The active area on Ceres is less than 1 km2, too small to
cause global albedo and spectral variations detectable in our data. Impact ejecta due to
impacting projectiles of tens of meters in size like those known to cause observable changes to
the surface albedo on Asteroid Scheila cannot cause detectable albedo change on Ceres due to its
relatively large size and strong gravity. The water vapor activity on Ceres is independent of
Ceres’ heliocentric distance, ruling out the possibility of comet-like sublimation process as a
possible mechanism driving the activity.
Keywords: minor planets, asteroids: individual (1 Ceres) – methods: observational – techniques:
photometric – technique: imaging processing – space vehicles
2
1. Introduction
Observations of dwarf planet Ceres by NASA’s Dawn spacecraft have revealed a surface
peppered with bright spots associated with impact craters (Nathues et al. 2015). The bright spots
are prominent albedo features on the otherwise uniform surface of Ceres in albedo and color.
Compositional analysis reveals that these bright spots are likely made up of a mixture of water
ice, salts, and dark background material, suggesting a briny subsurface source. Bright spots
within impact craters Dantu (125 km) and Occator (92 km) have been linked to intermittent and
localized water vapor sources observed by the Herschel Space Observatory in 2011-2013
(Küppers et al. 2014), and therefore possibly related to the sublimation activity. Recent ground-
based observations of Ceres also suggested short-term global spectral variability that was
attributed to changing amount of water ice on the surface (Perna et al. 2015). These observations
suggest a possibly changing surface caused by water outgassing on time scales of decades to
months. We searched for evidence of surface albedo and spectral variations that might be
evidence for ongoing activity by comparing various historical datasets and the most recent Dawn
Framing Camera (FC) data collected during the approach to Ceres. Our search covered three
different time frames: long-term in decades; mid-term in years, and short-term in months.
2. Analysis
2.1 Full-disk spectral variability
The spectral data archived in the Planetary Data System Small Bodies Node from four
epochs (Vilas et al. 1998, Bus & Binzel 2003, Lazzaro et al. 2006), from Perna et al. (2015) in
two epochs, and from recent ground-based observations (Reddy et al. 2015a) span over nearly
three decades, suitable for searching for long-term spectral variability of Ceres (Fig. 1a). The
3
spectra of Ceres from various datasets all show a nearly linear shape between 0.53 and 0.85 µm,
with increasing spectral slopes with solar phase angle, and absorptions <0.54 µm and >0.9 µm.
Therefore we fit linear slopes to all spectra between 0.54 and 0.85 µm to quantify the spectral
shape variation. Fig. 1b suggests that the variations in slope are ascribable to variations in phase
angle, except for the Dawn RC2 observations at 44º phase angle (described below), possibly
indicating that the approximate linear relationship between spectral slope and phase angle stops
somewhere between 20º and 40º phase angle.
Dawn FC observed Ceres for at least one full rotation of Ceres of 9.075 hours
Chamberlain et al. 2007) in three separate epochs in February and April 2015 during the
approach. These three observations are named Rotational Characterization (RC1, RC2, and
RC3, see Table 1). The disk of Ceres did not fill the field of view of the FC during the RCs.
Image calibration and spectral cube computation have been performed according to Reddy et al.
(2012a) and Nathues et al. (2014). The time difference between RC1 and RC2 is one week and
between RC2 and RC3 ten weeks, enabling us to explore short-term (days/weeks/months) global
spectral variability on Ceres. To minimize dependence on the effect of the absolute calibration
of the instrument, we calculated the relative spectra in each of the three RCs. The phase angle of
RC3 is about 8º. Based on the linear fit to the spectral slope with respect to phase angle (Fig.
1b), the spectral slope of Ceres at the RC3 phase angle is near zero. Therefore, the ratio spectra
of RC1 and RC2 to RC3 should remove the effect of the uncertain absolute radiometric
calibration of Dawn FC (Schröder et al. 2013a, 2014), and approximate the spectra of Ceres
collected at RC1 (17º phase angle) and RC2 (44º). The spectral slope from RC1 is consistent
with the overall trend with phase angle in the three RCs. We therefore based our spectral slope
analysis of Dawn FC data on these relative spectra (Fig. 1a).
4
2.2 Global photometric variability
To measure global variations in the albedo of Ceres over a time scale of years, we used
previously published HST data collected in 2003 and 2004 with the Advanced Camera for
Survey (ACS) High Resolution Channel (HRC) through the broadband F555W filter (Li et al.
2006), supplemented by observations taken with the Wide Field Camera 3 (WFC3) UVIS
channel through the broadband filter F555W in 2014 (GO-13503, PI: B.E. Schmidt). The
WFC3/UVIS F555W has a similar combined throughput as the ACS/HRC F555W filter in 2014.
We scaled the total flux of Ceres to a phase angle of 6º by a linear fit to the disk-integrated phase
function measured from the HST ACS/HRC data, which span a small range of phase angles from
5.4º to 7.7º, and phasing each flux measurement to the rotation of Ceres based on the rotational
period of Ceres, which has an accuracy of 7 ms per rotation (Chamberlain et al. 2007). The total
reflected flux of Ceres measured through the three datasets agrees with each other within 3%,
well within the absolute photometric calibration uncertainties of the two instruments.
2.3 Occator bright spots
The HST/ACS/HRC data and Dawn FC data are separated by 12 years, and are both
spatially resolved, suitable for searching for albedo changes in local scales on Ceres in a medium
timeframe of years. We focus on the Occator crater because the bright spots inside this crater are
undoubtedly the most prominent features on Ceres, and possibly related to the water vapor
activity (Nathues et al. 2015). Comparison of absolute feature brightness between different
datasets is extremely difficult, due primarily to the vastly different optical characteristics and
absolute calibrations between HST and FC. Our approach is to compare the Occator bright spots
5
with a reference region, for which we chose the Haulani crater (32 km diameter), because it is
relatively easy to identify in HST images. The underlying assumptions for our comparative
approach is that the reflectances of Occator bright spots and Haulani crater did not change
simultaneously, and this is justified by the unchanging disk-integrated brightness of Ceres over
the same time period (§2.2). We used RC1 images and RC3-equator images (Table 1) in this
study. Note that Haulani crater is in fact the brightest feature on Ceres in HST images (Li et al.
2006, Thomas et al. 2005), brighter than Occator crater (Fig. 2). The present study seeks an
answer to the question: Did the brightness of Occator crater change since the previous HST
observations, or are the different relative reflectance values of Haulani and Occator an observing
artifact from comparing the HST and Dawn datasets?
2.3.1 Photometric modeling
Precise photometric comparisons require corrections for different illumination and
viewing geometries, which further rely on photometric models of the whole surface of Ceres and
the bright spots. We separately modeled the photometric properties of Occator bright spots and
the global average of Ceres based on RC3 clear filter (effective wavelength 730 nm, FWHM 370
nm) data at 1.3 km/pixel with the Hapke model (Hapke 2012), following the procedure outlined
in Li et al. (2013). The shape model derived from Dawn FC data collected during approach is
used, coupled with the spacecraft trajectory and pointing data in NASA’s Navigation and
Ancillary Information Facility SPICE kernels, to calculate the local illumination and
scattering geometry. Some details of the photometric modeling are available in Li et al. (2015).
Due to the limited observing geometry available in our data, especially the lack of data at
small phase angles to constrain the opposition effect, some modeled photometric parameters for
6
Occator bright spots, such as the single-scattering albedo (SSA) and geometric albedo, highly
depend on the assumptions in the modeling as well as the particular form of photometric models
used. The Bond albedo, on the other hand, is much less model-dependent because it is
dominated by the light scattered towards moderate range of phase angles (30º-90º) that is well
sampled in the data.
Despite relatively large uncertainties in some modeled parameters, photometric models
show that the bright spots have significantly different properties from the average Ceres. The
SSA of the brightest part of Occator bright spots is between 0.67 and 0.80, comparing to the
average SSA of Ceres of 0.094-0.11. Occator bright spots have a Bond albedo of 0.24±0.01
through the Dawn FC clear filter, comparing to 0.034±0.001 for Ceres average. Since SSA and
Bond albedo measure the scattered light towards all directions, they are more fundamental
physical properties than geometric albedo, which is often poorly constrained due to the lack of
data at low phase angle. Hence we compared Occator bright spots with other solar system
objects in terms of their Bond albedos. The Bond albedo of Occator bright spots is higher than
the average Vesta and most asteroids in the main belt, but is comparable to or slightly lower than
the bright material on Vesta (Li et al. 2012) (Fig. 3). Compared to icy moons in the outer solar
system, Occator brightest spot is comparable or brighter than Ganymede and Callisto, but darker
than Europa and most large Saturnian satellites whose surfaces are dominated by water ice.
Therefore, Occator bright spots are unlikely pure water ice, but either admixtures of water ice,
salts, and perhaps other bright minerals with darker background material.
The bright spots appear to scatter light more isotropically, with a single-term Henyey-
Greenstein parameter of -0.1±0.05, than the average Ceres surface (-0.35±0.05), which is
consistent with most asteroids (C- and S-types). The phase function of a surface is affected by,
7
among other factors, grain transparency, shape, defects, impurities, etc. The relatively less
backscattering phase function of the bright spots is consistent with higher albedo and stronger
multiple scattering, and possibly more transparent particles. The Hapke model roughness
parameter for the bright spots is between 40º and 50º, significantly higher than the average Ceres
of 20º±3º. Relatively higher roughness suggests more loosely packed regolith particles that cast
substantial shadows compared to the average Ceres, consistent with a geologically young
regolith that has not been substantially processed (Schröder et al. 2013b).
2.3.2 Albedo variability of Occator bright spots
For comparisons of the brightness between Occator bright spots and Haulani crater in
different datasets, the effects of different resolution and the vastly different point-spread function
(PSF) have to be accounted for. From the radiometrically calibrated images of RC1 and RC3,
first, we downsampled them to the pixel scale of HST images (30 km/pixel), then convolve the
images with a PSF of HST ACS/HRC at F555W filter generated by the TinyTIM (Krist et al.
2011). We then measured the intensity value of the brightest pixel that contains the Occator
crater and Haulani crater from both HST images and the processed Dawn FC images. After
correcting for the varying local illumination and viewing angles for both features in all
measurements using the best-fit photometric parameters of Ceres surface, we plot the brightness
measurements with respect to local solar time for both features and compare their relative
brightness (Fig. 2b). The effect of subpixel shift in the downsampling process is negligible.
Ideally, after the correction for local illumination and viewing angle with the best-fit
photometric model, the brightness of both features should be independent of local solar time.
The residual variations with local solar time as shown in Fig. 2b suggest that the photometric
8
model does not completely remove the dependence on (varying) local scattering geometry over a
Ceres day for both features. The slightly different illumination and observing geometries in
different datasets result in different trends as they rotate over the disk of Ceres. However, note
that the variations after correction in Fig. 2b are at the 4% level, within the performance
expectation of photometric correction in general.
Fig. 2b shows that the reflectance contrast between the Occator region and the Haulani
region remains unchanged within 2% in HST, and Dawn FC RC1 and RC3 images. The reason
that Occator crater appeared to be darker than Haulani region in HST images, as well as in the
processed RC1 and RC3 images (Fig. 2b), is due to the large footprint of HST pixels and the
wide PSF (80% energy encircled at 5 pixel; Maybhate et al. 2010) of the ACS/HRC camera,
together washing out the high contrast of the relatively small bright spots (<4 km) with respect to
the background. The bright Haulani crater region, together with the bright rays outside of the
crater, are >50 km across. Fig. 2b also shows that no brightness variation was observed in
Occator between the RC1 and RC3 observations over the span of eleven weeks. We will
continue monitoring the brightness of the bright spots as well as all possible active regions on
Ceres throughout the Dawn mission at resolutions down to 35 m/pixel to search for any possible
change.
3. Discussions
To interpret the putative albedo variability on Ceres we calculated the area in the Occator
bright spots and how much albedo change is needed to create a measurable change in global
albedo of Ceres. Our albedo variation investigation is limited by the HST pixel scale of 30 km.
The Occator bright spots have a combined area of about 95 km2, accounting for 11% of the HST
9
pixel footprint. Therefore our study suggests that any brightness change within the bright spots,
or change in their total area, is smaller than 15%.
Given the water production rate of 2×1026 molecules s-1 for Ceres (Küppers et al. 2014),
assuming a Bond albedo of 0.034, and a thermal emissivity of 0.9, pure water ice sublimation
model (Cowan & A’Hearn 1979) suggests an active area of 0.2 km2 for a slow rotator case where
the surface is at instantaneous thermal equilibrium with local solar insolation at the latitude of
Occator (~20º). The low thermal inertia of Ceres of <15 J m-2 s-0.5 K-1 (Chamberlain et al. 2009)
justifies the slow rotator assumption. This active area corresponds to 0.2% of the combined
surface area of the Occator crater spots, or 0.5% of the central core bright spot. The active area
within the bright spots is too small to produce measurable albedo changes at 30 km pixel size.
The impact ejecta observed on main belt asteroid (596) Scheila during the impact in 2010
(Ishiguro et al. 2011) produced an average brightness change of ~5% over 10,000 km2 area. If
we use the crater size scaling law (Housen & Schmid 1983) with an exponent parameter of 0.7,
based on an escape velocity of 55 m s-1 for Scheila (as calculated by Ishiguro et al. 2011) and 0.5
km s-1 for Ceres, the same impact on Ceres could produce an ejecta field that is 104 smaller, or
~1 km2. Scaling with albedo and size shows that such an impact is still too small to produce any
albedo change on Ceres detectable by our analysis. Based on Bodewits et al. (2011)
consideration, the projectile on Scheila had a size of 35-60 m and a relative velocity of ~5.3 km
s-1; and the impact frequency is once every 103 years. Therefore, if the outgassing on Ceres were
triggered by sporadic impact events similar to that considered for Scheila, then such events
would not directly produce albedo changes with ejecta near the impact site detectable here.
Perna et al. (2015) suggested that Ceres displays short-term (months) visible spectral
variability, and attributed it to “extended resurfacing processes such as cryovolcanism or
10
cometary activity”. Our interpretation of the ground-based visible wavelength spectral data
shows an increasing spectral slope with increasing phase angle (Fig. 1). Observation geometry
has been shown to change slope and band depth in reflectance spectra of small bodies (Sanchez
et al. 2012), including Vesta (Reddy et al. 2012, Li et al. 2013). Hence our observations do not
support the results reported in that study. The current intermittent and weak water sublimation
does not cause albedo and color variations on the surface of Ceres detectable in our data.
Our analysis of the heliocentric dependence of activity on Ceres (Fig. 4) shows no
correlation between the two. This is an indication that whatever mechanism responsible for
water vapor outgassing on Ceres is at least not completely driven by solar insolation as for
comets. The presence of a network of cracks (Reddy et al. 2015b) inside craters with bright spots
such as Occator hints at an internal source that is responsible for the transportation of volatiles
from their subsurface sources.
We thank the Dawn operations team for the development, cruise, orbital insertion, and
operations of the Dawn spacecraft at Ceres. The Framing Camera project is financially
supported by the Max Planck Society and the German Space Agency, DLR. Li is supported by a
subcontract from the University of California, Los Angeles under the NASA Contract
#NNM05AA86 Dawn Discovery Mission.
This research made use of Astropy (Astropy
Collaboration et al. 2013), and Matplotlib (Hunter 2007)
Figure Captions
11
Figure 1. a) The visible spectra of Ceres taken at increasing phase angles (marked at the left of
each spectrum) from bottom to top (vertically shifted for clarity). From bottom to top: Geo:
Modeled geometric albedo spectrum from ground-based photometric observations (Reddy et al.
2015); P01, P02, P03: Perna et al. (2015) spectra on December 18 and 19, 2012; V87: Vilas et al.
(1998) spectrum in 1987; P04, P05, P06: Perna et al. (2015) spectra on January 18, 19, 20, 2013;
SOS: S2OS3 (Lazzaro et al. 2006); RC1: Dawn RC1; SII: SMASS II (Bus & Binzel 2003); V92:
Vilas et al. (1998) spectrum in 1992; and RC2: Dawn RC2. b) The linear spectral slopes of
Ceres’ spectra between 0.54 and 0.85 µm. The colors of symbols correspond to the colors in
panel a. The large error bar for the red symbol at 0.8º phase angle, corresponding to P02
spectrum, is due to the much higher noise in the spectrum compared to others. The black dashed
line is a linear fit to the spectral slope with respect to phase angle, not including the slope
measured from Dawn RC2 at 44º phase angle. The spectral slope of Ceres increases nearly
linearly from 0º to 25º phase angle.
Figure 2. a) Images of Haulani crater (left column) and Occator crater (right column) from HST
taken in December 2003 and January 2004 (top row), Dawn RC1 images taken on February 4,
2015 (second row), Dawn RC3 images taken on May 5, 2015 (third row), and the downsampled
and convolved images of Dawn RC3 images (bottom row). The to craters in all images are
marked by the dashed-line circles, and their names are noted in the RC3 images. The HST
images shown in the top row are drizzled (Fruchter & Sosey 2009) to 0”.015 per pixel scale from
the original 0”.025. The pixel scale at Ceres shown in the top row is 18 km for HST images, for
RC1 images 13.7 km, for RC3 images 1.32 km, and for processed RC3 images 30 km. SC and
SS are the longitude and latitude of, respectively, sub-spacecraft/HST and sub-solar points; and
12
ph is the phase angle. b) The photometrically corrected reflectance of Haulani region (red) and
Occator region (blue) in three datasets. The reflectance are measured from original HST images,
and Dawn RC1 and RC3 images that are convolved with HST/ACS/HRC PSF and downsampled
to HST image resolution. The I/F are corrected to a common, arbitrary geometry of 3.15º
incidence angle, 3º emission angle, and 6.15º phase angle, close to the phase angle of HST
images. The horizontal lines are the averages of the I/F for both features to show the contrast
between the two regions in various cases. The contrast between Haulani region and Occator
region at 30 km scale is 5-7%, and remain unchanged within 3% between HST, RC1, and RC3.
Figure 3. Comparisons of Bond albedos between Ceres (red square), Occator bright spots (blue
square), and other solar system objects (green circles). The Bond albedos of icy satellites are
from a list complied by Verbiscer et al. (2012); and those of comets (9P, 19P, 81P, and 103P),
asteroids (asteroid numbers 4, 243, 253, 433, 951, 2867, and 26143), the Moon, and the Martian
satellites are from a list compiled by Li et al. (2013).
Figure 4. The positions of Ceres in its orbit during the previous searches of water vapor. The
solid ellipse is Ceres’ orbit with the Sun at the center, and “Peri” marks the perihelion of Ceres.
The filled symbols mark the detection of water vapor (circles, by Herschel) or OH (triangles, by
IUE), and the open symbols are non-detection. The directions of the triangles mark the direction
of off-limb OH search, where upper triangles are off northern limb, and down triangles are off
southern limb. The S and N are summer solstice for southern hemisphere and northern
hemisphere, respectively, based on the pole measurement from Dawn data. The blue shaded part
of Ceres’ orbit is during Dawn’s planned stay from April 1st, 2015 to June 30th, 2016. No
13
obvious trend of water outgassing and orbital phase (perihelion/aphelion, seasons) is clearly
identified. The outgassing appears to be intermittent and occurs at arbitrary orbital phases.
14
Table 1. Observation conditions extracted from the FC color data.
Phase
Average
S/C
distance
Heliocentric
Angle (deg)
resolution
(km)
distance (AU)
(km/pixel)
17.2 – 17.6
7.68
42.7 – 45.3
4.24
RC1
RC2
RC3 (Equatorial) 7.7 – 11.0
1.28
82,151
45,359
13,603
2.8529
2.8578
2.9051
15
References
Astropy Collaboration, Robitaille, T.P., Tollerud, E.J., et al. 2013, A&A, 558, A33
Bodewits, D., Kelley, M.S., Li, J.-Y., et al. 2011, ApJL, 773, L3
Bus, S. & Binzel, R.P. Small Main-belt Asteroid Spectroscopic Survey, Phase II. EAR-A-I0028-
4-SBN0001/SMASSII-V1.0. NASA Planetary Data System (2003)
Chamberlain, M., Sykes, M.V., & Esquerdo, G.A. 2007, Icar, 188, 451
Chamberlain, M.A., Lovell, A.J., & Sykes, M.V. 2009, Icar, 202, 487
Cowan, J.J. & A’Hearn, M.F. 1979, The Moon and the Planets, 21, 155
De Sanctis, M.C., Ammannito, E., Raponi, A., et al. 2015, Nature, 528, 241
Fruchter, A. & Sosey, M. 2009. The Multidrizzle Handbook, Version 3.0, (Baltimore, STScI)
Hapke, B. Theory of Reflectance and Emittance Spectroscopy, 2nd Ed. Cambridge Univ. Press,
Cambridge, UK.
Housen, K.R. & Schmidt, R.M. 1983, J. Geophys. Res., 88, 2485
Hunter, J.D. 2007, Computing in Science & Engineering, 9, 90
Ishiguro, M., Hanayama, H., Hasegawa, S., et al. 2011, Astrophys. J., 740, L11
Krist, J.E., Hook, R.N., Stoehr, F. 2011, Proc. of SPIE, 8127, 81270J-1
Lazzaro D. et al. Small Solar System Objects Spectroscopic Survey V1.0. EAR-A-I0052-8-
S3OS2-V1.0. NASA Planetary Data System (2006)
Li, J.-Y., McFadden, L.A., Parker, J.Wm., et al. 2006, Icar, 182, 143
Li, J.-Y., Mittlefehldt, D.W., Pieters, C.M., et al. 2012, LPICo, 1667, 6386
Li, J.-Y., Le Corre, L., Schröder, S.E., et al. 2013, Icar, 226, 1252
Li, J.-Y., Nathues, A., Le Corre, L., et al. 2015, EPSC2015-383
16
Maybhate, A. et al. 2010, ACS Instrument Handbook, Version 10.0 (Baltimore: STScI)
Nathues, A., Hoffmann, M., Cloutis, E.A., et al. 2014, Icar, 239, 222
Nathues, A., Hoffmann, M., Schaefer, M., et al. 2015, Nature, 528, 237
Küppers, M., O’Rourke, L., Bocklelée-Morvan, D., et al. 2014, Nature, 505, 525
Perna, D., Kaňuchová, Z., Ieva, S., et al. 2015, Astron. Astrophys., 575, L1
Reddy, V., Nathues, A., Le Corre, L., et al. 2012a, Science, 336, 700
Reddy, V., Sanchez, J.A., Nathues, A., et al. 2012b, Icar, 217, 153
Reddy, V., Li, J.-Y., Gary, B.L., et al. 2015a, Icar, 260, 332
Reddy, V., Nathues, A., Le Corre, L., et al. 2015b, MetSoc 2015, 5161
Sanchez, J.A., Reddy, V., Nathues, A., et al. 2012, Icar, 220, 36
Schröder, S.E., Maue, T., Gutiérrez Marqués, P., et al. 2013a, Icar, 226, 1304
Schröder, S.E., Mottola, S., Keller, H.U. 2013b, Planet. Space Sci., 85, 198
Schröder, S.E., Mottola, S., Matz, K.-D., et al. 2014, Icar, 234, 99
Thomas, P.C., Parker, J.Wm., et al. 2005, Nature, 437, 224
Verbiscer, A.J., Helfenstein, P., & Buratti, B.J. 2012, The Science of Solar System Ices, 47
Vilas, F. et al. Vilas Asteroid Spectra V1.1, EAR-A-3-RDR-VILAS-ASTEROID-SPECTRA-
V1.1. NASA Planetary Data System (1998)
17
|
1512.05440 | 2 | 1512 | 2016-02-26T10:03:37 | ALMA Observations of a Gap and a Ring in the Protoplanetary Disk around TW Hya | [
"astro-ph.EP"
] | We report the first detection of a gap and a ring in 336 GHz dust continuum emission from the protoplanetary disk around TW Hya, using the Atacama Large Millimeter/Submillimeter Array (ALMA). The gap and ring are located at around 25 and 41 au from the central star, respectively, and are associated with the CO snow line at ~30 au. The gap has a radial width of less than 15 au and a mass deficit of more than 23%, taking into account that the observations are limited to an angular resolution of ~15 au. In addition, the 13CO and C18O J = 3 - 2 lines show a decrement in CO line emission throughout the disk, down to ~10 au, indicating a freeze-out of gas-phase CO onto grain surfaces and possible subsequent surface reactions to form larger molecules. The observed gap could be caused by gravitational interaction between the disk gas and a planet with a mass less than super-Neptune (2M_{Neptune}), or could be the result of the destruction of large dust aggregates due to the sintering of CO ice. | astro-ph.EP | astro-ph |
Draft version June 13, 2018
Preprint typeset using LATEX style emulateapj v. 5/2/11
ALMA OBSERVATIONS OF A GAP AND A RING IN THE PROTOPLANETARY DISK AROUND TW HYA
Hideko Nomura1,*, Takashi Tsukagoshi2,*, Ryohei Kawabe3,4,5, Daiki Ishimoto6,1, Satoshi Okuzumi1, Takayuki
Muto7, Kazuhiro D. Kanagawa8, Shigeru Ida9, Catherine Walsh10, T.J. Millar11 and Xue-Ning Bai12
Draft version June 13, 2018
ABSTRACT
We report the first detection of a gap and a ring in 336GHz dust continuum emission from the proto-
planetary disk around TW Hya, using the Atacama Large Millimeter/Submillimeter Array (ALMA).
The gap and ring are located at around 25 and 41 AU from the central star, respectively, and are
associated with the CO snow line at ∼ 30AU. The gap has a radial width of less than 15AU and a
mass deficit of more than 23%, taking into account that the observations are limited to an angular
resolution of ∼ 15AU. In addition, the 13CO and C18O J = 3 − 2 lines show a decrement in CO
line emission throughout the disk, down to ∼ 10AU, indicating a freeze-out of gas-phase CO onto
grain surfaces and possible subsequent surface reactions to form larger molecules. The observed gap
could be caused by gravitational interaction between the disk gas and a planet with a mass less than
super-Neptune (2MNeptune), or could be the result of the destruction of large dust aggregates due to
the sintering of CO ice.
Subject headings: protoplanetary disks -- stars:
individual (TW Hya) -- submillimeter: planetary
systems -- planet-disk interactions -- molecular processes
1. INTRODUCTION
The physical structures and chemical compositions of
gas, dust, and ice in protoplanetary disks control the for-
mation process of planets and the composition of their
cores and atmospheres. The ALMA long baseline cam-
paign has detected gaps and rings in dust continuum
emission from the circumstellar disk around a very young
star (∼ 0.1 − 1Myrs), HL Tau (ALMA Partnership et al.
2015). The origin of this complex disk structure in such
a young object remains under debate.
In this Letter, we present ALMA observations of the
relatively old (∼ 3 − 10Myrs) gas-rich disk around the
young Sun-like star, TW Hya (∼ 0.8M⊙), which show
[email protected]
1 Department of Earth and Planetary Sciences, Tokyo Insti-
tute of Technology, 2-12-1 Ookayama, Meguro, Tokyo, 152-8551,
Japan
2 College of Science, Ibaraki University, Mito, Ibaraki 310-
8512, Japan
3 National Astronomical Observatory of Japan, 2-21-1 Osawa,
Mitaka, Tokyo 181-8588, Japan
4 SOKENDAI (The Graduate University for Advanced Stud-
ies), 2-21-1 Osawa, Mitaka, Tokyo 181-8588, Japan
5 Department of Astronomy, School of Science, University of
Tokyo, Bunkyo, Tokyo 113-0033, Japan
6 Department of Astronomy, Graduate School of Science,
Kyoto University, Kitashirakawa-Oiwake-cho, Sakyo-ku, Kyoto
606-8502, Japan
7 Division of Liberal Arts, Kogakuin University, 1-24-2
Nishi-Shinjuku, Shinjuku-ku, Tokyo, 163-8677, Japan
8 Institute of Low Temperature Science, Hokkaido University,
Sapporo 060-0819, Japan
9 Earth-Life Science Institute, Tokyo Institute of Technology,
2-12-1 Ookayama, Meguro, Tokyo 152-8550, Japan
10 Leiden Observatory, Leiden University, P. O. Box 9513,
2300 RA Leiden, The Netherlands
11 Astrophysics Research Centre, School of Mathematics and
Physics, Queen's University Belfast, University Road, Belfast
BT7 1NN, UK
12 Institute
for Theory
Smithsonian Center
MS-51, Cambridge, MA 02138, USA
for Astrophysics,
and Computation, Harvard-
60 Garden Street,
∗
These authors contributed equally to this work.
the presence of a gap and a ring associated with the
CO snow line. TW Hya's proximity (54±6 pc) makes
it an ideal source to study formation environment of
a planetary system (e.g., Andrews et al. 2012). The
disk is old compared with other gas-rich protoplane-
tary disks whose lifetime is typically ∼ 3Myrs (e.g.,
Hern´andez et al. 2007). Nevertheless, the disk gas mass
is > 0.05M⊙ inferred through HD line observations with
Herschel (Bergin et al. 2013).
In exoplanetary systems
giant planets have been discovered at/beyond Neptune-
orbit around Sun-like stars by direct imaging obser-
vations using Subaru/HiCIAO (Thalmann et al. 2009;
Kuzuhara et al. 2013). Thus, planets are able to form
even at large distances within disk lifetime by, for exam-
ple, scattering of planetary cores (Kikuchi et al. 2014).
Given that planet-disk interaction is key for planet or-
bital evolution and planet population synthesis (e.g.,
Kley & Nelson 2012; Ida et al. 2013), understanding gap
formation in protoplanetary disks helps to gain insight
into the early evolution of our own Solar System, as well
as the observed diversity of exoplanetary systems.
2. OBSERVATIONS AND DATA REDUCTION
TW Hya was observed with ALMA in Band 7
on 20-21 May, 2015 with 40 antennas in Cycle 2
with a uv-coverage of 22-580 kλ (PI: D. Ishimoto).
The spectral windows were centered at 329.295GHz
(SPW1), 330.552GHz (SPW0), 340.211GHz (SPW2),
and 342.846GHz (SPW3), covering C18O J = 3 − 2,
13CO J = 3 − 2, CN N = 3 − 2 and CS J = 7 − 6.
The channel spacing was δν =30.52kHz and the band-
width was 117.188MHz except for SPW 2 in which a
channel spacing of δν =15.26kHz and a bandwidth of
58.594MHz were used. The quasar J1058+0133 was ob-
served as a bandpass calibrator while the nearby quasar
J1037-2934 was used for phase and gain calibration. The
mean flux density of J1037-2934 was 0.58Jy during the
observation period.
The visibility data were reduced and calibrated using
2
Nomura et al.
CASA, version 4.3.1 and 4.4.0. The visibility data was
reduced for each SPW separately, and the continuum vis-
ibilities were extracted by averaging the line-free chan-
nels in all SPWs. The corrected visibilities were imaged
using the CLEAN algorithm with Briggs weighting with
a robust parameter of 0 after calibration of the bandpass,
gain in amplitude and phase, and absolute flux scaling,
and then flagging for aberrant data. In addition to the
usual CLEAN imaging, we performed self-calibration of
the continuum emission to improve the sensitivity and
image quality. The obtained solution table of the self-
calibration for the continuum emission was applied to
the visibilities of the lines. The self-calibration signifi-
cantly improved the sensitivity of the continuum image
by one order of magnitude from 2.6 to 0.23mJy. In ad-
dition, the continuum visibility data with deprojected
baselines longer than 200kλ was extracted in order to
enhance small scale structure around the gap. The high
spatial resolution data was analysed in a similar way,
but imaged using the CLEAN algorithm with uniform
weighting. Also, the ALMA archived data of N2H+
J = 4 − 3 line at 372.672GHz were reanalysed in a sim-
ilar way to compare with that obtained by our observa-
tions. The re-analysed data is consistent with that in
Qi et al. (2013). For the N2H+ line data, the synthe-
sized beam and the 1σ RMS noise level in 0.1 km s−1
were 0."44×0."41 (P.A. = 12.4◦) and 31.2mJy beam−1.
3. RESULTS
3.1. Dust Continuum Emission
The observed dust continuum emission maps at
336GHz are plotted for both the full data (Fig. 1a)
and high spatial frequencies only (> 200kλ, Fig. 1b).
The synthesized beam and RMS noise were 0."37×0."31
(∼ 20AU, P.A. = 55.7◦) and 0.23mJy beam−1. The total
and peak flux density were 1.41Jy and 0.15Jy beam−1
with a SN (signal-to-noise) ratio of 705 and 652, re-
spectively. The results agree well with the previous
SMA observations (Andrews et al. 2012) and ALMA ob-
servations (Hogerheijde et al. 2015) with lower spatial
resolution and sensitivity. The synthesized beam and
RMS for the data at high spatial resolution data only
were 0."32×0."26 (∼ 15AU, P.A. = 54.5◦) and 0.49mJy
beam−1.
As shown by the black line in Fig. 2a, the dust con-
tinuum emission of the full data has a shallow dip in
the radial profile of the surface brightness obtained by
deprojecting the observed image data assuming an incli-
nation angle of 7◦ and a position angle of −30◦ (Qi et al.
2008). By imaging the data at high spatial frequen-
cies only (> 200kλ), we identify a gap and a ring at
around 25AU and 41AU, respectively (Fig. 1b and gray
line in Fig. 2a). The uv-coverage of the full data set
(from 22 to 580 kλ) is sufficiently high and indeed vital
for this analysis. The continuum visibility profile as a
function of the deprojected baseline length is plotted in
Fig. 3. Our result is consistent with the recent ALMA
observations with higher spatial resolution and sensitiv-
ity (Zhang et al. 2016). Since the high spatial frequency
data misses flux of spatially extended structure, the to-
tal flux is lower than that of the full data by an order of
magnitude. Artificial structure appears at R > 70AU in
the high spatial frequency data probably due to a fail-
Fig. 1. -- 336GHz dust continuum map imaged using (A) all spa-
tial frequency data and (B) only high spatial frequency (> 200kλ)
data, (C) 13CO J = 3 − 2 and (D) C18O J = 3 − 2 integrated line
emission maps. Emission less than 5σ-noise-level were masked.
The synthesized beam sizes are shown in the bottom left corner of
each panel.
ure to subtract the side lobe pattern of the synthesized
beam.
Its intensity is less than the 2σ-noise-level and
the structure is masked in Fig. 2a. The location of the
gap is similar to that of the axisymmetric depression in
polarized intensity of near-infrared scattered light imag-
ing of dust grains recently found by Subaru/HiCIAO and
Gemini/GPI (Akiyama et al. 2015; Rapson et al. 2016).
3.2. Dust Surface Density and Gap Parameters
We extract the radial dust surface density distribution
(Fig. 2b), using the deprojected data of the dust contin-
uum emission and the equation,
Id,ν = Bν(Td)[1 − exp(−τd,ν)],
(1)
where Id,ν is the observed intensity of the dust contin-
uum emission at the frequency ν, Bν(Td) is the Planck
function at the dust temperature, Td. The optical depth,
τd,ν, is defined as τd,ν = κνΣdust, where Σdust is the
dust surface density and the dust opacity, κν , is set as
κν = 3.4 cm2 g−1. We adopt Td = 22K(R/10AU)−0.3,
where R is the disk radius, by fitting the model result in
Andrews et al. (2012) and assume a uniform temperature
distribution in the vertical direction. The derived dust
optical depth is shown in Fig. 2c. We note that the de-
rived dust surface density distribution is consistent with
that derived from SMA observations of dust continuum
emission and model calculations (Andrews et al. 2012).
Also, the assumed dust temperature is consistent with
the color temperature obtained from the spectral index of
continuum emission across the observed four basebands
between 329GHz and 343GHz (Fig. 2d) in the optically
thick regions (≤ 30AU, Fig. 2c).
To estimate the gap width, ∆gap, and depth, (Σ0 −
Σ)/Σ0, where Σ0 and Σ are the basic surface density
and the surface density with the gap at the gap center,
A Gap and a ring in the TW Hya disk
3
102
101
]
1
-
m
a
e
b
y
J
m
[
ν
I
100
(A)
0
10
]
2
-
m
c
g
[
10-1
d
Σ
150
100
50
0.4
0.3
0.2
0.1
10 20 30 40
τ
1.2
1.0
0.8
0.6
0.4
0.2
0.0
3.0
(C)
10 20 30 40
α
2.0
(B)
-2
10
0
10 20 30 40 50 60 70 80
Radius [AU]
(D)
0 10 20 30 40 50 60 70 80
1.0
Radius [AU]
0 = 1.1 × 103R−1.5
Fig. 2. -- Radial distributions of (A) intensity of 336GHz dust
continuum emission, (B) dust surface density, (C) optical depth
and (D) spectral index between 329GHz and 343GHz. Gray line in
Figure (A) shows the data of only high spatial frequency with its
Gaussian fit (green line). Red crosses in (A)-(C) are derived from
high spatial frequency data recovered by the full spatial frequency
data at the gap. The inserts show a close-up of the region around
the gap.
the intensity at the gap, Igap (red crosses in Fig. 2a), is
derived by completing the missing flux of the high spatial
frequency data, I high
gap (gray line in Fig. 2a), with the full
spatial frequency data, I full (black line in Fig. 2a). The
intensity profile across the gap was obtained as follows;
(i) fit the high frequency data without the gap using a
power-law profile of I high
AU mJy beam−1,
(ii) fit the gap region in the high frequency data using
a single Gaussian profile of ∆Igap = I high
gap =
7.7 exp(−(RAU − 24.7AU)2/{2(5.9AU)2}) mJy beam−1
(whose FWHM is 13.9±1.0AU and depth is 7.7±0.5mJy)
(green line in Fig. 2a) and (iii) obtain the intensity at the
gap as Igap = I full − ∆Igap, where I full is the intensity of
the full data. In this fitting, we adopted the edges of the
gap at R = 9 − 13AU (inner edge) and 39 − 43 AU (outer
edge) so that ∆Igap is overlaid to I full smoothly. The op-
tical depth at the gap is derived using Igap and Eq.(1),
and then the dust surface density at the gap is derived
using the same method mentioned above (red crosses in
Fig. 2b). The resulting gap width is ∆gap ∼ 15AU and
the depth is (Σ0 − Σ)/Σ0 ∼ 0.23. We note that our esti-
mate of the gap width is limited by the angular resolution
of the high spatial frequency data. The pattern of the
gap and ring is also affected by residual artefacts due to
the cut-off at 200kλ, which introduces uncertainties in
the location of the gap center, and the width and depth
of the gap.
− I high
0
3.3. CO Isotopologue Line Emission
The observed 13CO and C18O J = 3 − 2 rotational line
emission maps are plotted in Figs. 1c and 1d. The result-
ing synthesized beam and the 1σ RMS noise level in 0.03
km s−1 width-channels were 0."41×0."33 (P.A. = 53.4◦),
12.0mJy beam−1 (13CO) and 0."41×0."33 (P.A. =
53.8◦), 13.4 mJy beam−1 (C18O). The integrated inten-
sity maps were created by integrating from 1.18 to 4.30
km s−1 (13CO) and from 1.48 to 4.15 km s−1 (C18O). The
resultant noise levels of the map were 8.4 mJy beam−1
km s−1 (13CO) and 6.5 mJy beam−1 km s−1 (C18O). The
deprojected radial profiles of the integrated line emission
of the 13CO and C18O following subtraction of the dust
]
y
J
[
l
a
e
R
1.4
1.2
1.0
0.8
0.6
0.4
0.2
0.0
−0.2
0
]
y
J
[
y
r
a
n
g
a
m
i
I
0.06
0.05
0.04
0.03
0.02
0.01
0.00
−0.01
−0.02
0
100
200
300
400
500
600
deprojected uv distance [k-lambda]
100
200
300
400
500
600
deprojec ed uv dis ance [k-lambda]
Fig. 3. -- Real (top) and imaginary (bottom) parts of the az-
imuthally averaged continuum visibility profile as a function of the
deprojected baseline length.
continuum emission are plotted in Fig. 4a. The signal-
to-noise ratio is not sufficiently high to analyse the high
spatial resolution data only.
3.4. CO Column Density
We obtain the CO radial column density distribution,
using the deprojected data of the 13CO and C18O line
observations (Fig. 4d). The CO column density, NCO,
is derived from the integrated line emission of the C18O
J = 3 − 2 (Turner 1991; Rybicki & Lightman 1979), as-
suming the abundance ratios of CO :13CO=67 : 1 and
13CO : C18O=7 : 1 (Qi et al. 2011). The optical depth
of the C18O line emission, τC18O,v (Fig. 4c), is obtained
from the observed line intensity as follows. We assume
that the CO line emitting region is mainly in the sur-
face layer, while the dust continuum emitting region is
near the midplane. Thus, the deprojected intensity of
the CO line emission plus dust continuum emission can
be approximately derived from the following equation by
simply assuming three zones in the vertical direction of
the disk,
IC18O+cont,v = Bν(Tg)(1 − e−τC18O,v /2)
+Bν(Td)(1 − e−τd,ν )e−τC18O,v /2
(2)
+Bν(Tg)(1 − e−τC18O,v /2)e−τd,ν −τC18O,v /2,
where Tg is the gas temperature of the CO line emitting
region and Td is the dust temperature near the midplane.
The data continuum-subtracted data, IC18O+cont,v −Id,ν,
together with Eq.(1) is used for the analysis. Since
the observed line ratio of R IC18O,vdv/ R I13CO,vdv is
larger than 1/7 (Fig. 4b), the 13CO line is optically
thick. Therefore, we adopt the brightness temperature
at the peak of the 13CO line (Fig. 5) as the gas tem-
perature, Tg. Also, we adopt a dust temperature of
Td = 22K(R/10AU)−0.3 and the dust optical depth,
τd,ν, is derived from the dust continuum observations
(Fig. 2c). We note that in the analysis we simply assume
that the gas temperature is uniform in the vertical direc-
tion, and the gas temperature of the C18O line emitting
region is the same as that of the 13CO line emitting re-
gion, although it could be lower in reality. In addition,
C18O could be depleted due to the effect of isotopologue-
dependent photodissociation, and NCO/NC18O is possi-
4
Nomura et al.
)
O
10-1
8
1
C
(
e
v
a
τ
10-2
(C)
]
2
-
m
c
[
)
1018
O
C
N
(
1017
(D)
0
10 20 30 40 50 60 70 80
Radius [AU]
(
R
H
2
g
/
d
)
1
0.1
-3
10
)
d
/
g
O
C
R
(
-4
10
(E)
0 10 20 30 40 50 60 70 80
Radius [AU]
Fig. 4. -- Radial distributions of (A) integrated line emission of
13CO (red line) and C18O (blue line) J = 3 − 2, (B) line ratio
of C18O to 13CO, (C) averaged optical depth of the C18O line,
(D) CO column density and (E) CO gas-to-dust surface density
ratio. H2 gas-to-dust ratio is also marked for a reference by simply
assuming the abundance ratio of CO to H2 of 6 × 10−5. All data
are smoothed to a beam size of 0."45×0."45.
bly higher than 440 (Miotello et al. 2014). The CO col-
umn density will be underestimated due to these effects.
3.5. CO Gas-to-Dust Surface Density Ratio
Using the obtained CO column density and dust sur-
face density (Figs. 4d and 2b), we derive the CO gas-to-
dust surface density ratio (Fig. 4e). If we convert it to the
H2 surface density, assuming an abundance ratio of CO
to H2 of 6 × 10−5 (Qi et al. 2011), the estimated H2 gas
mass is orders of magnitude lower than that predicted
from the observations of the HD line emission by the
Herschel Space Observatory (Bergin et al. 2013). The
resulting H2 gas-to-dust surface density ratio (∼ 0.1 − 1)
is about two orders of magnitude lower than the typi-
cal interstellar value of ∼100, which suggests strong CO
depletion throughout the disk down to ∼ 10AU.
The CO gas-to-dust surface density ratio increases be-
yond a radius of ∼ 40AU since the dust surface den-
sity drops dramatically in this region. It could be due
to drift of pebbles from the outer disk toward the cen-
tral star (e.g, Takeuchi et al. 2005; Andrews et al. 2012;
Walsh et al. 2014b).
4. DISCUSSIONS
4.1. Origin of the Gap and Ring in the Dust
Continuum Emission
The gap and ring resemble those in the HL Tau sys-
tem, recently found by the ALMA long baseline cam-
paign (ALMA Partnership et al. 2015). Our result shows
that gaps and rings in the (sub)millimeter dust con-
tinuum can exist, not only in relatively young disks
(0.1 − 1Myrs) but also in relatively old disks (3 −
10Myrs). One possible mechanism to open a gap is
the gravitational interaction between a planet and the
gas (e.g., Lin & Papaloizou 1979; Goldreich & Tremaine
1980; Fung et al. 2014). Such an interaction may also
produce the spiral density waves recently found in optical
and near-infrared scattered light imaging of dust grains
in protoplanetary disks (e.g., Muto et al. 2012). Accord-
ing to recent theoretical analyses of gap structure around
a planet (Kanagawa et al. 2015a,b, 2016), the depth and
width of the gap are controlled by the planetary mass,
the turbulent viscosity and the gas temperature. The
shape of the gap is strongly influenced by angular mo-
mentum transfer via turbulent viscosity and/or instabil-
ity caused by a steep pressure gradient at the edges of a
gap. The observed gap has an apparent width and depth
of ∆gap ∼ 15AU and (Σ0 − Σ)/Σ0 ∼ 0.23, respectively.
This is too shallow and too wide compared with that
predicted by theory. However, the observations are lim-
ited to an angular resolution of ∼ 15AU, and the depth
and width could be deeper and narrower in reality. For
instance, if we assume that the gap depth times the gap
width retains the value derived from the observations,
it is possible for the gap to have a width and depth of
∆gap ∼6AU and (Σ0 − Σ)/Σ0 ∼ 0.58, which is similar to
the GPI result (Rapson et al. 2016). Such a gap could be
opened by a super-Neptune-mass planet, depending on
parameters of the disk, such as the turbulent viscosity
(Kanagawa et al. 2015a,b, 2016). If the gap in the larger
dust grains is deeper than that in the gas, the planet
could be lighter than super-Neptune mass. We note that
a planet of even a few Earth masses, although it can-
not open a gap in the gas, can open a gap in the dust
distribution if a certain amount of pebble-sized particles,
whose motion is not perfectly coupled to that of gas, are
scattered by the planet and/or the spiral density waves
excited by the planet (Paardekooper & Mellema 2006;
Muto & Inutsuka 2009).
Another possible mechanism to form a gap and an as-
sociated ring in dust continuum emission is the micro-
scopic process of sintering of CO ice on dust aggregates
(Okuzumi et al. 2016; Sirono 2011). The gaps and rings
observed in the younger and more luminous HL Tau sys-
tem could be explained by the sintering of various molec-
ular ices in the disk at their distinct snow line locations
(Okuzumi et al. 2016). Although our observations in-
dicate that the CO depleted region is located down to
∼ 10AU in the TW Hya system, model calculations of
the temperature in the disk suggest the CO snow line is
located at ∼ 30AU. Sintering is a process that renders
an aggregate less sticky (Sirono 2011). Just outside the
CO snow line where sintering occurs, large aggregates be-
come easily destroyed into small fragments by collisions
and their radial drift motion by the above-mentioned
mechanism slows down. Thus dust grains are stuck just
outside the CO snow line, and a bright ring and a dark
lane inside the ring is formed in the dust continuum emis-
sion. If another sintering region is formed inside the dark
lane by, for example, CH4 sintering, the region between
the bright CO and CH4 sintering regions would look like
a gap. According to model calculations (Okuzumi et al.
2016), the CH4 sintering region is located at ∼ 10AU in
the TW Hya disk.
4.2. CO Gas Depletion inside the CO Snow Line
A Gap and a ring in the TW Hya disk
5
40
30
]
K
20
[
)
k
a
e
p
(
B
T
10
9
8
7
10
13
CO w/ cont
CN w/ cont
CN w/o cont
+
N2H
w/ cont
+
w/o cont
N2H
30
50
70
90
Radius [AU]
Fig. 5. -- Radial distributions of the observed brightness temper-
ature at the peak of 13CO J = 3 − 2 (red) and N2H+ J = 4 − 3
(green) lines. The dust temperature at the midplane (dot-dashed
gray line) and the brightness temperature of CN N = 3−2 (purple)
line are also plotted for comparison. Dashed lines show the data
subtracted by the dust continuum emission for N2H+ and CN. All
data are smoothed to a beam size of 0."45×0."45.
From the CO line observations, we find a very low
column density of CO compared with dust throughout
the disk down to a radius of about 10AU. This CO de-
pletion could indicate a general absence of H2 gas com-
pared with dust. However, the Herschel HD observations
(Bergin et al. 2013) indicates that it is more likely that
it is due to CO freeze-out on grains with the possibil-
ity that subsequent grain surface reactions form larger
molecules even inside the CO snow line (Favre et al.
2013; Williams & Best 2014). The CO column density
derived from our observations, NCO ∼ 1018cm−2, can be
explained by detailed model calculations using a chem-
ical network which includes freeze-out of molecules on
grains and grain-surface reactions (Aikawa et al. 2015).
The model calculations show the CO depletion will pro-
ceed inside the CO snow line due to the sink effect: con-
version of CO to less volatile species on grain surface.
In the case of TW Hya, it could occur on a timescale
shorter than the disk lifetime, depending on the amount
of small grains. See Aikawa et al. (2015) for more de-
tailed discussions including the effect on other species.
The CO depletion spread over the disk is inconsistent
with the prediction by the previous ALMA N2H+ ob-
servations that depletion would be localized beyond the
CO snow line (Qi et al. 2013). This could be because the
N2H+ line emission traces the disk surface and not the
CO depleted region.
Fig. 5 shows the brightness temperature at the peak
of the 13CO J = 3 − 2 line obtained by our obser-
vations and the N2H+ J = 4 − 3 line at 372.672GHz
obtained by the ALMA archived data (2011.0.00340.S).
Since the 13CO line is optically thick, the brightness tem-
perature at the peak of the line emission represents the
gas temperature of the line emitting region. If LTE is
applicable, the N2H+ brightness temperature is higher
than the gas temperature of the 13CO line emitting re-
gion at the disk radius of ∼ 40AU, and higher than the
dust temperature near the midplane down to ∼ 15AU
(Fig. 5). If the N2H+ line is optically thin, the gas tem-
perature of the line emitting region is higher than the
brightness temperature. Therefore, the N2H+ line should
come from the surface layer of the disk. Model calcula-
tions also predict that N2H+ exists in the disk surface
for the model with (sub)micron-sized grains, similar to
radical species abundant in the disk surface, such as CN
and C2H (e.g., Walsh et al. 2010; Aikawa et al. 2015).
Our results (Fig. 5) and the SMA observations of C2H
(Kastner et al. 2015) show that the radial intensity pro-
files of these species are similar. They have peaks around
the disk radius of 40AU beyond which the dust surface
density drops. The result suggests that in order to trace
the CO depleted region, the C18O line may be more ro-
bust than the N2H+ line.
In the CO depleted region complex organic molecules
would be produced via grain surface reactions since
hydrogen attachment
to pro-
duce methanol and more
(e.g.,
Watanabe & Kouchi 2008; Walsh et al. 2014a). Methyl
cyanide, which has recently been detected from a proto-
planetary disk for the first time by ALMA ( Oberg et al.
2015), could be formed through such grain surface reac-
tions.
complex species
to CO is
thought
We would like to thank the referee for his/her com-
ments which improved our paper. We are also grate-
ful to Sean Andrews and Joel Kastner for their fruit-
ful comments. This paper makes use of the follow-
ing ALMA data: ADS/JAO.ALMA#2013.1.01397.S and
ADS/JAO.ALMA#2011.0.00340.S. ALMA is a partner-
ship of ESO (representing its member states), NSF
(USA) and NINS (Japan), together with NRC (Canada),
NSC and ASIAA (Taiwan) and KASI (Republic of
Korea),
in cooperation with the Republic of Chile.
The Joint ALMA Observatory is operated by ESO,
AUI/NRAO and NAOJ. This work is partially supported
by Grants-in-Aid for Scientific Research, 23103005,
25108004, 25400229 and 15H03646. T.T. was supported
by JSPS KAKENHI grant No. 23103004. C.W. is sup-
ported by the Netherlands Organisation for Scientific Re-
search (program number 639.041.335). Astrophysics at
QUB is supported by a grant from the STFC.
Facility: ALMA.
Aikawa, Y., Furuya, K., Nomura, H., & Qi, C. 2015, ApJ, 807, 120
Akiyama, E., Muto, T., Kusakabe, N. et al. 2015, ApJ, 802, L17
ALMA Partnership, Brogan, C. L., P´erez, L. M. et al. 2015, ApJ,
808, L3
Andrews, S. M., Wilner, D. J., Hughes, A. M. et al. 2012, ApJ,
744, 162
Bergin, E. A., Cleeves, L. I., Gorti, U. et al. 2013, Nature, 493,
644
Favre, C., Cleeves, L. I., Bergin, E. A., Qi, C., & Blake, G. A.
2013, ApJ, 776, L38
Fung, J., Shi, J.-M., & Chiang, E. 2014, ApJ, 782, 88
Goldreich, P., & Tremaine, S. 1980, ApJ, 241, 425
REFERENCES
6
Nomura et al.
Hern´andez, J., Hartmann, L., Megeath, T., et al. 2007, ApJ, 662,
Okuzumi, S., Momose, M., Sirono, S.-i., Kobayashi, H., &
1067
Hogerheijde, M. R., Bekkers, D., Pinilla, P., Salinas, V. N.,
Kama, M., Andrews, S. M., Qi, C., & Wilner, D. J. 2016, A&A,
589, A99
Ida, S., Lin, D. N. C., & Nagasawa, M. 2013, ApJ, 775, 42
Kanagawa, K. D., Muto, T., Tanaka, H., Tanigawa, T., Takeuchi,
T., Tsukagoshi, T., & Momose, M. 2015a, ApJ, 806, L15
-- . 2016, PASJ, submitted
Kanagawa, K. D., Tanaka, H., Muto, T., Tanigawa, T., &
Takeuchi, T. 2015b, MNRAS, 448, 994
Kastner, J. H., Qi, C., Gorti, U., Hily-Blant, P., Oberg, K.,
Forveille, T., Andrews, S., & Wilner, D. 2015, ApJ, 806, 75
Kikuchi, A., Higuchi, A., & Ida, S. 2014, ApJ, 797, 1
Kley, W., & Nelson, R. P. 2012, ARA&A, 50, 211
Kuzuhara, M., Tamura, M., Kudo, T., et al. 2013, ApJ, 774, 11
Lin, D. N. C., & Papaloizou, J. 1979, MNRAS, 186, 799
Miotello, A., Bruderer, S., & van Dishoeck, E. F. 2014, A&A,
572, A96
Muto, T., & Inutsuka, S.-i. 2009, ApJ, 695, 1132
Muto, T., Grady, C. A., Hashimoto, J., et al. 2012, ApJ, 748, L22
Oberg, K. I., Guzm´an, V. V., Furuya, K., Qi, C., Aikawa, Y.,
Andrews, S. M., Loomis, R., & Wilner, D. J. 2015, Nature, 520,
198
Tanaka, H. 2016, ApJ, in press (astro-ph/1510.03556)
Paardekooper, S.-J., & Mellema, G. 2006, A&A, 453, 1129
Qi, C., D'Alessio, P., Oberg, K. I., Wilner, D. J., Hughes, A. M.,
Andrews, S. M., & Ayala, S. 2011, ApJ, 740, 84
Qi, C., Wilner, D. J., Aikawa, Y., Blake, G. A., & Hogerheijde,
M. R. 2008, ApJ, 681, 1396
Qi, C., Oberg, K. I., Wilner, D. J., et al. 2013, Science, 341, 630
Rapson, V. A., Kastner, J. H., Millar-Blanchaer, M. A., & Dong,
R. 2015, ApJ, 815, L26
Rybicki, G. B., & Lightman, A. P. 1979, Radiative processes in
astrophysics (New York: Wiley)
Sirono, S.-i. 2011, ApJ, 735, 131
Takeuchi, T., Clarke, C. J., & Lin, D. N. C. 2005, ApJ, 627, 286
Thalmann, C., Carson, J., Janson, M., et al. 2009, ApJ, 707, L123
Turner, B. E. 1991, ApJS, 76, 617
Walsh, C., Millar, T. J., & Nomura, H. 2010, ApJ, 722, 1607
Walsh, C., Millar, T. J., Nomura, H., Herbst, E., Widicus
Weaver, S., Aikawa, Y., Laas, J. C., & Vasyunin, A. I. 2014a,
A&A, 563, A33
Walsh, C., Juh´asz, A., Pinilla, P., et al. 2014b, ApJ, 791, L6
Watanabe, N., & Kouchi, A. 2008, Progress In Surface Science,
83, 439
Williams, J. P., & Best, W. M. J. 2014, ApJ, 788, 59
Zhang, K., Bergin, E. A., Blake, G. A., Cleeves, L. I., Hogerheijde,
M., Salinas, V., & Schwarz, K. R. 2016, ApJ, 818, L16
|
1204.1468 | 1 | 1204 | 2012-04-06T12:44:48 | Anelastic tidal dissipation in multi-layer planets | [
"astro-ph.EP"
] | Earth-like planets have viscoelastic mantles, whereas giant planets may have viscoelastic cores. The tidal dissipation of such solid regions, gravitationally perturbed by a companion body, highly depends on their rheology and on the tidal frequency. Therefore, modelling tidal interactions presents a high interest to provide constraints on planets' properties and to understand their history and their evolution, in our Solar System or in exoplanetary systems. We examine the equilibrium tide in the anelastic parts of a planet whatever the rheology, taking into account the presence of a fluid envelope of constant density. We show how to obtain the different Love numbers that describe its tidal deformation. Thus, we discuss how the tidal dissipation in solid parts depends on the planet's internal structure and rheology. Finally, we show how the results may be implemented to describe the dynamical evolution of planetary systems. The first manifestation of the tide is to distort the shape of the planet adiabatically along the line of centers. Then, the response potential of the body to the tidal potential defines the complex Love numbers whose real part corresponds to the purely adiabatic elastic deformation, while its imaginary part accounts for dissipation. This dissipation is responsible for the imaginary part of the disturbing function, which is implemented in the dynamical evolution equations, from which we derive the characteristic evolution times. The rate at which the system evolves depends on the physical properties of tidal dissipation, and specifically on how the shear modulus varies with tidal frequency, on the radius and also the rheological properties of the solid core. The quantification of the tidal dissipation in solid cores of giant planets reveals a possible high dissipation which may compete with dissipation in fluid layers. | astro-ph.EP | astro-ph | Astronomy&Astrophysicsmanuscript no. RMZL12author3bisaccptarxiv
November 5, 2018
© ESO 2018
Anelastic tidal dissipation in multi-layer planets
F. Remus1,2,3, S. Mathis3,4, J.-P. Zahn1, V. Lainey2
1 LUTH, Observatoire de Paris -- CNRS -- Universit´e Paris Diderot, 5 place Jules Janssen, F-92195 Meudon Cedex, France
2 IMCCE, Observatoire de Paris -- UMR 8028 du CNRS -- Universit´e Pierre et Marie Curie, 77 avenue Denfert-Rochereau, F-75014
3 Laboratoire AIM Paris-Saclay, CEA/DSM -- CNRS -- Universit´e Paris Diderot, IRFU/SAp Centre de Saclay, F-91191 Gif-sur-
Paris, France
Yvette, France
4 LESIA, Observatoire de Paris -- CNRS -- Universit´e Paris Diderot -- Universit´e Pierre et Marie Curie, 5 place Jules Janssen,
F-92195 Meudon, France
e-mail: [email protected], [email protected], [email protected], [email protected]
Received December 6, 2011; accepted February 24, 2012
ABSTRACT
Context.Earth-like planets have viscoelastic mantles, whereas giant planets may have viscoelastic cores. The tidal dissipation of such
solid regions, gravitationally perturbed by a companion body, highly depends on their rheology and on the tidal frequency. Therefore,
modelling tidal interactions presents a high interest to provide constraints on planets' properties and to understand their history and
their evolution, in our Solar Sytem or in exoplanetary systems.
Aims. The purpose of this paper is to examine the equilibrium tide in the anelastic parts of a planet whatever the rheology, taking into
account the presence of a fluid envelope of constant density. We show how to obtain the different Love numbers that describe its tidal
deformation. Thus, we discuss how the tidal dissipation in solid parts depends on the planet's internal structure and rheology. Finally,
we show how the results may be implemented to describe the dynamical evolution of planetary systems.
Methods. We expand in Fourier series the tidal potential, exerted by a point mass companion, and express the dynamical equations
in the orbital reference frame. The results are cast in the form of a complex disturbing function, which may be implemented directly
in the equations governing the dynamical evolution of the system.
Results. The first manifestation of the tide is to distort the shape of the planet adiabatically along the line of centers. Then, the
response potential of the body to the tidal potential defines the complex Love numbers whose real part corresponds to the purely
adiabatic elastic deformation, while its imaginary part accounts for dissipation. The tidal kinetic energy is dissipated into heat through
anelastic friction, which is modeled here by the imaginary part of the complex shear modulus. This dissipation is responsible for
the imaginary part of the disturbing function, which is implemented in the dynamical evolution equations, from which we derive the
characteristic evolution times.
Conclusions. The rate at which the system evolves depends on the physical properties of tidal dissipation, and specifically on how
the shear modulus varies with tidal frequency, on the radius and also the rheological properties of the solid core. The quantification
of the tidal dissipation in solid cores of giant planets reveals a possible high dissipation which may compete with dissipation in fluid
layers.
Key words. Planet-Star interactions -- Planets and satellites: general -- Planets and satellites: physical evolution -- Planets and
satellites: dynamical evolution and stability
1. Introduction and general context
Since 1995 a large number of extrasolar planets have been dis-
covered, which display a wide diversity of physical parameters
(Santos & et al., 2007). Quite naturally the question arose of their
habitability, i.e. whether they could allow the development of
life. Determining factors are the presence of liquid water and of
a protective magnetic field, properties which are closely linked
to the values of the rotational and orbital elements of the plane-
tary systems. And these elements strongly depend on the action
of tides. Once a planetary system is formed in a turbulent accre-
tion disk, its fate is determined by the initial conditions and the
mass ratio between planet and hosting star. Through tidal inter-
action between components, the system evolves either to a stable
state of minimum energy (where all spins are aligned, the orbits
are circular and the rotation of each body is synchronized with
Send offprint requests to: F. Remus
the orbital motion) or the companion tends to spiral into the par-
ent body. Indeed, by converting kinetic energy into heat through
internal friction, tidal interactions modify the orbital and rota-
tional properties of the components of the considered system,
and thus their structure through internal heating. This mecha-
nism depends sensitively on the internal structure and dynamics
of the perturbed body.
Recent studies have been carried out on tidal effects in fluid bod-
ies like stars and envelopes of giant planets (Ogilvie & Lin 2004-
2007; Ogilvie 2009; Remus, Mathis & Zahn 2012). However
the planetary solid regions, such as mantles of Earth-like plan-
ets or rocky cores of giant planets, if present (e.g. Guillot 1999,
Gaulme et al. 2011), may contribute likewise to tidal dissipa-
tion. The first study of a tidally deformed elastic body was done
by Lord Kelvin (1863) who applied it to an incompressible ho-
mogeneous Earth. Further developments were made by Love
(1911), who introduced a set of dimensionless numbers, the so-
called Love numbers, to quantify the tidal perturbation. More
1
2
1
0
2
r
p
A
6
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
8
6
4
1
.
4
0
2
1
:
v
i
X
r
a
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
recently Greff-Lefftz (2005) generalized these results in the case
of a spheroidal rotating Earth. In the meanwhile Dermott (1979)
considered a two-layer model and he studied the impact of a
tidally deformed static fluid shell on the deformation of an elas-
tic solid core.
If the body is not perfectly elastic, i.e. if its internal structure
is anelastic, the tidal deformation presents a lag with respect to
the tension exerted by the tidal force, and causes the dissipation
of kinetic energy. Several recent studies addressed this prob-
lem of tidal dissipation using linear viscoelastic models. Peale
& Cassen (1978) evaluated the tidal dissipation in the Moon
considering various models of internal structure. Tobie et al.
(2005) applied the Maxwell rheological model to evaluate the
dissipation in Titan and Europa; Ross & Schubert (1986) com-
pared three different linear models of viscoelasticity (Kelvin-
Voigt, Maxwell, Standard Anelastic Solid), to which Henning
et al. (2009) added the Burgers body. All these studies suscitate
our interest in the tidal dissipation resulting from the anelastic
deformation of the solid parts of a planet when perturbed by a
companion.
We shall study here the tidal dissipation in a planet which pos-
sesses an anelastic core, made of a mix of ice and rock, sur-
rounded by a fluid envelope, such as an ocean. The planet is
part of a binary system where what we call the companion (or
perturber) may be either the hosting star or a satellite of the
planet. Due to the tide exerted by the companion, the core of
the two-layer planet is deformed elastically, but because of the
anelasticity of the material composing the core, this deformation
is accompanied by a viscous dissipation that we evaluate what-
ever the rheology. As an illustration, the results will be given
for a Maxwell body. We then compare the value of the tidal dis-
sipation in presence of a fluid envelope with that achieved by
the fully solid planet, and we examine the dependence of the re-
sults on the relative sizes of the core and the planet, the relative
densities, and the viscoelastic parameters. In the last section, we
establish the equations governing the dynamical evolution of the
system, from which we deduce the caracteristic times of circu-
larization, synchronization and spin alignments.
2. Elastic deformations of a solid body under tidal
perturbation
2.1. Thesystem
We consider a planet A of mass MA, consisting of a rocky (or icy)
core and a fluid envelope, rotating at the angular velocity Ω and
tidally perturbed by a second body of mass MB, assumed to be
ponctual, moving around A on a Keplerian orbit, of semi-major
axis a and eccentricity e, at the mean motion n. We locate any
point M in space by its usual spherical coordinates (r, θ, ϕ) in
a spin equatorial reference frame RE : {A, XE, YE, ZE} centered
on body A and whose axis (A, ZE) has the direction of the rota-
tion axis of A. The corresponding configuration is illustrated on
Figure 1.
In this section, we first assume that planet A has no fluid layer
and its internal structure is supposed to be perfectly elastic. We
will then denote by ρ its density and R its mean radius.
2.2. Thetidalpotential
The planet is submitted to a tidal force, exerted by the perturber
B, which derives from a perturbing time dependent potential
2
Fig. 1. Spherical coordinates system attached to the equatorial
reference frame RE : {A, XE, YE, ZE} associated to body A. A
point M is located by r ≡ (r, θ, ϕ); the point mass body B by
rB ≡ (rB, θB, ϕB).
l (θ, ϕ) in RE.
U (r, t). Following Zahn (1966a-b) and generalizing it by using
Kaula (1962), Lambeck (1980), Yoder (1995-1997) and Mathis
& Le Poncin (2009) (hereafter MLP09) in the present case of a
close binary system where spins are not aligned, the components
are not synchronized with the orbital motion and where the or-
bit is not circular, we expand the tidal potential U in spherical
harmonics Ym
Before we proceed, we need to define the Euler angles
link the spin equatorial frame RE : {A, XE, YE, ZE} of
that
the central body A, on one hand, and the orbital frame
RO : {A, XO, YO, ZO}, on the other hand, to the quasi-inertial
frame RR : {A, XR, YR, ZR} whose axis ZR has the direction of
the total angular momentum of the whole system.
We need the three following Euler angles to locate the orbital
reference frame RO with respect to RR:
-- I, the inclination of the orbital plane of B;
-- ω∗, the argument of the orbit pericenter;
-- Ω∗, the longitude of the orbit ascending node.
The equatorial reference frame RE is defined by three other Euler
angles with respect to RR:
-- ε, the obliquity of the rotation axis of A;
-- Θ∗, the mean sideral angle defined by Ω = dΘ∗/dt;
-- φ∗, the general precession angle.
Refer to Figure 2 for an illustration of the relative position of
these three reference frames and the associated angles. For con-
venience, all the following developpements will be done in the
spin equatorial frame RE of A (as it has been done in MLP09).
All following results are derived from the Kaula's transform
(Kaula 1962), used to explicitly express the whole generic multi-
pole expansion in spherical harmonics of the perturbing potential
U in terms of the Keplerian orbital elements of B in the equato-
rial A-frame:
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
j m d2
j,m (ε)
2
2
2
1
1
0
2
1
0
1
0
0
(cid:16)cos ε
2(cid:17)4
2(cid:17)3(cid:16)sin ε
2(cid:17)
−2(cid:16)cos ε
√6(cid:16)cos ε
2(cid:17)2(cid:16)sin ε
2(cid:17)2
2(cid:17)2(cid:16)sin ε
− 3(cid:16)cos ε
(cid:16)cos ε
2(cid:17)4
2(cid:17)2
√6 cos ε(cid:16)cos ε
2(cid:17)(cid:16)sin ε
2(cid:17)
1 − 6(cid:16)cos ε
2(cid:17)2(cid:16)sin ε
2(cid:17)2
−
Table 1. Values of the obliquity function d2
where j > m obtained from Eq. (8), (from MLP09).
j,m (ε) in the case
axis of A, as:
a3
d2
(7)
Um, j,p,q(r) = (−1)m s (2 − m)! (2 − j)!
(2 + m)! (2 + j)!
hd2
j,m(ε) F2, j,p(I) G2,p,q(e)i r2 ,
× G MB
j,m(γ) is defined, for j > m, by:
where G is the gravitational constant.
The obliquity function d2
(2 + m)!(2 − m)!# 1
j,m(γ) = (−1) j−m" (2 + j)!(2 − j)!
2(cid:19)(cid:21) j−m
2(cid:19)(cid:21) j+m (cid:20)sin(cid:18) γ
×(cid:20)cos(cid:18) γ
where Pα,β
(x) are the Jacobi polynomials (cf. MLP09). The val-
ues of these functions, for indices j < m, are deduced from
d2
j,m(π + γ) = (−1)2− jd2
− j,m(γ) or from their symmetry proper-
− j,−m(γ) = d2
j,m(γ) = (−1) j−md2
ties: d2
m, j(−γ); moreover, we have:
d2
j,m(0) = δ j,m. Values are given in Table 1.
We also define, the inclination function F2, j,p(I):
P( j−m, j+m)
2− j
(cos γ) ,
(8)
l
2
F2, j,p(I) = (−1) j
(2 + j)!
4 p! (2 − p)!
2(cid:19)(cid:21) j−2p+2 (cid:20)sin(cid:18) I
×(cid:20)cos(cid:18) I
with the symmetry property:
F2,− j,p(I) ="(−1)2− j (2 − j)!
2− j
2(cid:19)(cid:21) j+2p−2
× P( j+2p−2, j−2p+2)
(2 + j)!# F2, j,p (I) .
(cos I) ,
(9)
(10)
Values are given in Table 2.
The eccentricity functions G2,p,q(e) are polynomial functions
having eq for argument (see Kaula, 1962). Their values for the
usual sets {2, p, q} are given in Table 3, knowing that in the case
of weakly eccentric orbits, the summation over a small number
of values of q is sufficient (q ∈ [ − 2, 2]). In the following, let us
denote by I = [ − 2, 2] × [ − 2, 2] × [0, 2] × Z the set in which
the quadruple {m, j, p, q} takes its values.
If we simplify the expansion of the potential in the case where
spins are aligned and perpendicular to the orbital plan, where
obliquity ε and orbital inclination I are zero, Eq. (5) reduces to
the expression of the potential given by Zahn (1977).
The tidal force induces a displacement of each particule consti-
tuting the planet, thus causing some deformations. In particular,
the core's surface is deformed as described by the Love theory
(Love 1911).
3
Fig. 2. Inertial reference (RR), orbital (RO), and equatorial (RE)
rotating frames, and associated Eulers angles of orientation.
1
rl+1
B
Pm
l (cos θB) eimϕB =
1
al+1
(l − j)!
(l + j)!
4π
lXp=0Xq∈Z
s 2l + 1
lXj=−l
j,m(ε) Fl, j,p(I) Gl,p,q(e) eiiiΨl,m, j,p,q
(1)
×dl
where θB and ϕB are respectively the colatitude and the longitude
of the point mass perturber B, and where the phase argument is
given by:
Ψl,m, j,p,q(t) = σl,m,p,q(n, Ω) t + τl,m, j,p,q (ω∗, Ω∗, φ∗) .
(2)
We have defined here the tidal frequency:
σl,m,p,q(n, Ω) = (l − 2p + q) n − mΩ ,
and the phase τl,m, j,p,q:
τl,m, j,p,q = (l − 2p)ω∗ + j(Ω∗ − φ∗) + (l − m)
π
2
.
(3)
(4)
We study here binary systems close enough for the tidal interac-
tion to play a role, but we also consider that the companion is far
(or small) enough to be treated as a point mass. We then are al-
lowed to assume the quadrupolar approximation, where we only
keep the first mode of the potential, l = 2:
U(r, θ, ϕ, t) = Re
2Xm=−2
2Xj=−2
2Xp=0Xq∈Z
Um, j,p,q(r) Pm
2 (cos θ) eiii Φ2,m, j,p,q(ϕ,t)
(5)
where
Φ2,m, j,p,q(ϕ, t) = mϕ + Ψ2,m, j,p,q(t)
(6)
The functions Um, j,p,q (r, θ) may be expressed in terms of the
Keplerian elements (the semi-major axis a of the orbit, its eccen-
tricity e and its inclination I) and the obliquity ε of the rotation
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
j
0
0
0
1
1
1
2
2
2
p
0
1
2
0
1
2
0
1
2
3
2
F2, j,p (I)
8 sin2 I
3
4 sin2 I + 1
− 3
8 sin2 I
3
4 sin I (1 + cos I)
− 3
2 sin I cos I
− 3
4 sin I (1 − cos I)
4 (1 + cos I)2
2 sin2 I
4 (1 − cos I)2
3
3
3
Table 2. Values of the inclination function F2, j,p(I). Values for
indices j < 0 can be deduced from Eq. (10), (from MLP09).
p
0
0
0
0
0
1
1
q
-2
-1
0
1
2
-2
-1
p
2
2
2
2
2
1
1
1
q
2
1
0
-1
-2
2
1
0
G2,p,q(e)
7
17
0
− 1
2 e + · · ·
1 − 5
2 e2 + · · ·
2 e + · · ·
2 e2 + · · ·
4 e2 + · · ·
2 e + · · ·
(cid:16)1 − e2(cid:17)−3/2
3
9
Table 3. Values of the eccentricity function G2,p,q(e), (from
MLP09).
2.3. Dynamicalequationsforasolidbodyandtheirboundary
conditions
To describe the internal evolution of the main component A sub-
mitted to the perturbations induced by the tidal potential pre-
sented above, we use the Eulerian formalism (Dahlen et al.
1999). The system of equations, needed to follow the motion of
a particule, is composed by the Eulerian momentum (11a) and
mass (11b) conservation laws, and the Poisson equation (11c)
satisfied by the potential Φ of self-gravitation:
ρ
∂2s
∂t2
∂s
= ∇∇∇ · ¯¯σσσ + ρ∇∇∇ (Φ + U) ,
+ ∇∇∇ · ρ
Φ = −4πGρ ,
∂t! = 0 ,
∂ρ
∂t
∇∇∇2
(11a)
(11b)
(11c)
where s designates the displacement vector and ¯¯σσσ is the stress
tensor. We complete this system with the constitutive equation to
link the stress exerted on the body to the resulting deformation.
Assuming that tidal effect corresponds to a traction applied on
the body, with no rotational contribution, the deformation tensor
reduces to the strain tensor ¯¯ǫǫǫ:
¯¯ǫǫǫ =
1
2h∇∇∇s + (∇∇∇s)Ti ,
(11d)
4
T
designates the transposed tensor of ¯¯hhh. We then get a
where ¯¯hhh
relation linking the stress tensor ¯¯σσσ to the strain tensor ¯¯ǫǫǫ that ac-
counts for the rheology of the body, and that we represent by a
function Frh:
(11e)
¯¯σσσ = Frh(¯¯ǫǫǫ) .
To solve this system (11), we need to apply boundary condi-
tions to the five previous equations, assuming that there is no
displacement (12a) neither attraction (12b) at the center of mass
r = 0, the gravitational potential has to be continuous (12c) and
the Lagrangian traction has to vanish (12d) at the surface r = R:
sr=0 = 0 ,
(Φ + U)r=0 = 0 ,
[Φ]R+
R− = 0 ,
i.e. :
= 0 .
(cid:0)er · ¯¯σσσ(cid:1)(cid:12)(cid:12)(cid:12)r=R
(12a)
(12b)
" ∂Φ
∂r
+ 4π G ρ sr#R+
R−
= 0 ,
(12c)
(12d)
2.4. Linearizationofthesystem
Assuming that tidal effects, and thus the resulting elastic de-
formation, are small amplitude perturbations to the hydrostatic
equilbrium, we are allowed to linearize the system (11) and its
boundary conditions (12). To do so, we expand a scalar quantity
X as:
X (r, θ, ϕ, t) = X0(r) + X′ (r, θ, ϕ, t) ;
(13)
X0 designates the spherically symmetrical profile of X, and X′
represents the perturbation due to the tidal potential. The dis-
placement s and the tidal potential U are also considered as per-
turbations. Thus, correct to first order in s, we obtain the fol-
lowing form of system (11):
∂2s
∂t2
ρ0
= ∇∇∇ · ¯¯σσσ + ρ0 ∇∇∇(cid:0)Φ′ + U(cid:1) + ρ′ ∇∇∇Φ0 ,
ρ′ + ∇∇∇ · (ρ0 s) = 0 ,
Φ′ = −4πGρ′ ,
¯¯σσσ = K −
2
3
µ! tr(cid:0)¯¯ǫǫǫ(cid:1) ¯¯III + 2µ¯¯ǫǫǫ ,
∇∇∇2
(14a)
(14b)
(14c)
(14d)
where we made use of the Hooke's law (14d), which is a lin-
ear constitutive law that governs elastic materials as long as the
load does not exceed the material's elastic limit, in the case of
an isotropic material (i.e.: whose properties are independent of
direction in space). It means that strain is directly proportional
to stress, through the bulk modulus K and the shear modulus µ.
The reference state is drawn from an up-to-date planetary struc-
ture model. It is governed by the following Poisson equation and
the static momentum equation:
∇∇∇2
Φ0 = −4πGρ0 ,
∇∇∇P0 = ρ0 ∇∇∇Φ0 ,
(15a)
(15b)
where we made use of the following convention for the gravity:
g0 = ∇∇∇Φ0.
2.5. Analyticalsolutionsforanhomogeneousincompressible
body
To solve the linear system (14), we expand all scalar quan-
tities in spherical harmonics Ym
l (θ, ϕ). Moreover, as all vecto-
rial quantities that intervene in Eqs. (14a-14b) are poloidal, we
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
may expand them in the basis of vectorial spherical harmonics
l (θ, ϕ)i, where R refers to the radial part and S to the
spheroidal part of a given vector (Rieutord 1987, Mathis & Zahn
2005):
l (θ, ϕ), Sm
hRm
∀(l, m) ∈ N × [ − l, l] ,
Rm
l (θ, ϕ) = Ym
Sm
l (θ, ϕ) = ∇∇∇ShYm
where ∇∇∇S designates the horizontal gradient:
l (θ, ϕ) er ,
l (θ, ϕ)i ,
(16a)
(16b)
(17)
∇∇∇S =
∂ ·
∂θ
eθ +
1
sin θ
∂ ·
∂ϕ
eϕ .
We introduce six radial functions ym
(r) (Takeuchi & Saito
1972) to expand all quantities in spherical harmonics, at the
quadrupolar approximation (l = 2):
{1...6}
-- the displacement:
lutions, considering the expansion (18): ∀m ∈ [ − 2, 2],
ym
1 (r) =Xj,p,q
2 (r) =Xj,p,q"2µ
ym
+
3
3
k2
k2
4
3
1
2
1
2
R2 +
r2!
R2 −
R2 − r2! Um, j,p,q(r) ,
r R gs 8
r2 R gs 8
r2!
R gs 8
πGρ2 k2
− ρ(1 + k2)#Um, j,p,q(r) ,
r2! Um, j,p,q(r) ,
R2 −
(R2 − r2) Um, j,p,q(r) ,
r R gs 4
r R gs
5
6
3
ym
ym
ym
3
k2
k2
8µ
3
3 (r) =Xj,p,q
4 (r) =Xj,p,q
5 (r) =Xj,p,q
6 (r) =Xj,p,q" 2(1 + k2)
r R gs r2 −
+4πGρ0
k2
r
ym
(1 + k2) Um, j,p,q(r) ,
8
3
R2!# Um, j,p,q(r) ,
(19a)
(19b)
(19c)
(19d)
(19e)
(19f)
s(r, θ, ϕ, t) = X(m, j,p,q)∈Ihym
1 (r) Rm
2 (θ, ϕ)
+ym
3 (r) Sm
2 (θ, ϕ)i eiiiΨ2,m, j,p,q(t) ,
-- the total potential:
(18a)
where we have introduced the acceleration of gravity at the sur-
face gs and the second-order Love number k2. The latter com-
pares the perturbed part Φ′(R) of the self-gravitational potential
at the surface of a fully-solid planet of mean radius R, deformed
by tidal force, with the tidal perturbing potential U(R):
(U+Φ′)(r, θ, ϕ, t) = X(m, j,p,q)∈I
5 (r) Ym
ym
2 (θ, ϕ) eiiiΨ2,m, j,p,q(t), (18b)
-- the Lagrangian traction:
er · ¯¯σσσ(r, θ, ϕ, t) = X(m, j,p,q)∈Ihym
2 (r) Rm
2 (θ, ϕ)
+ym
4 (r) Sm
2 (θ, ϕ)i eiiiΨ2,m, j,p,q(t) ,
(18c)
-- the Lagrangian attraction (introduced to express the continu-
ity of the gradient of the potential):
∀m ∈ [ − 2, 2], ym
6 (r) =
d
drhym
5 (r)i − 4πGρ0ym
1 (r) .
(18d)
The linear system governing the radial functions ym
(r) is
given in Appendix. In the case of an incompressible (K → +∞)
and homogeneous body (µ, ρ0 = cst), the system (14) con-
strained by boundary conditions (12) admits the following so-
{1...6}
def
=
k2
Φ′(R)
U(R)
.
(20)
The expression of k2 is established in Sect. 3, for an ocean-free
planet (Eq. 61) or a two-layer planet (Eq. 63).
3. Modified elastic tidal theory in presence of a fluid
envelope
We now assume that planet A is not entirely solid, but has a fluid
envelope. We follow the method proposed by Dermott (1979) to
evaluate how the anelastic dissipation is modified by the pres-
ence of a fluid layer surrounding the solid region. The first step
consists in determining the behaviour of the elastic response in
this configuration. We denote by Rc (resp. Rp) the mean radius
of the solid core (resp. of the whole planet, including the height
of the fluid layer); ρc and ρo designate the density of the core
and the ocean, both assumed to be uniform, as a first step. More
generally, all quantities will be written with a "c" subscript when
evaluated at the core boundary, and with a "p" subscript if taken
at the surface of the planet. The evolution of the system is de-
scribed in the orbital frame RO : {A, XO, YO, ZO} centered on A
and comoving with the perturber B. We will use polar coordi-
nates (r, Θ) to locate a point P, where r is the distance to the
center of A, and Θ is the angle formed by the radial vector and
the line of centers.
5
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
Fig. 3. Left: tidal displacement s. Middle: equatorial slice of s. Right: meridional slice of s. The orange arrow indicates the direction
of the perturber B, the red one corresponds to the rotation axis of A. The two slices are planes of symmetry.
3.1. Verticaldeformationattheboundaryofthecore
In RO, the tidal potential takes the form (Dermott 1979)
P2(cos Θ) ,
U(r) = −ζ(r) g(r) P2(cos Θ) = −ζc gc
where we have introduced the tidal height
r2
R2
c
and the gravity
r ,
ζ(r) =
MB
M(r) (cid:18) r
a(cid:19)3
g(r) = GM(r)
,
r2
(21)
(22)
(23)
M(r) being the fraction of mass of the planet inside the radius r.
The expression of the tidal potential in the rotating frame of
B (Eq. 21) is linked to its expression in the equatorial inertial
frame (Eq. 5), through the Kaula's transform (Eq. 1). Indeed, the
Legendre polynomia summation formula
P2(cos Θ) = Re
2Xm=0
(2 − m)!
(2 + m)!
(2 − δ0,m) Pm
2 (cos θ) eiiimϕ
× Pm
2 (cos θB) e−iiimϕB
(24)
involves the term Pm
transformed following Eq. (1-2) to obtain Eq. (5).
2 (cos θB) e−iiimϕB in Eq. (21) that has to be
In this section we are interested in the mofication of Love num-
bers due to the presence of a fluid envelope on top of the solid
core. Thus, we will focus on the deformations of the core's
surface, and particularly on the vertical displacements. Love
(1911) proved that tidal deformations could be described by the
same harmonic function than the tidal potential which causes it.
Therefore the equations of the core and planet boundaries are
respectively of the form
rc ≡ Rc + sr(Rc) = Rc [1 + S 2 P2(cos Θ)] ,
rp = Rp [1 + T2 P2(cos Θ)] .
(25a)
(25b)
Thus, sr(Rc) = Rc S 2 P2(cos Θ) represents the radial displace-
ment at the core's boundary corresponding to the vertical tidal
deformation of amplitude S 2 P2(cos Θ). In 1909, Love defined
6
the number h2, as the ratio between the amplitude of the verti-
cal displacement at the surface of the planet and the equilibrium
tidal height (disturbing potential divided by undisturbed surface
gravity, both taken at the surface of the core) in the case of a
fully-solid planet. Solving the whole system of equations, he de-
termined its expression as
sr(Rc)
U(Rc)/gc ≡
Rc S 2
1 + ¯µ
(26)
5
2
ζc
def
=
h2
1
,
=
where ¯µ is called the effective rigidity, in the sense that it evalu-
ates the relative importance of elastic and gravitational forces:
¯µ =
19µ
2ρcgRc
.
(27)
In presence of the fluid envelope, the ratio between the amplitude
of the tidal surface vertical displacement and the tidal height will
be modulated by a multiplicative factor F, due to the additional
loading exerted by the tidally-deformed fluid layer. We may then
introduce a new notation hF
2 for the modified Love number in
presence of a fluid envelope:
def
=
hF
2
sr(Rc)
U(Rc)/gc ≡
Rc S 2
ζc
= F h2 = F ×
5/2
1 + ¯µ
.
(28)
We have now to express this factor in function of the parameters
of the system. To do so, we have to list all the forces acting on
the surface of the core. Before carrying out the study of these
forces, let us introduce a specific notation. All physical quanti-
ties X(r) will be separated in two terms: the first corresponds to
the constant part that does not depend on where the quantity is
calculated; the second one (called the "effective deforming" con-
tribution and denoted X′(r)) is a term proportional to the spheri-
cal surface harmonic P2 (see 24).
3.2. Gravitationalforcesactingonthesurfaceofthecore
The planet is not only subjected to the direct action of the tidal
potential U, but also to the self-gravitational potential Φ per-
turbed by the first. In calculating the latter, we have to consider
both contributions of the solid core and the fluid envelope, Φc
and Φo respectively.
At any point r of the core, Φc(r) corresponds to the internal po-
tential created by the core :
Φc(r) = −
gc
Rc 3R2
c − r2
2
+
3
5
r2 S 2 P2! .
(29)
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
At the same point r, Φo(r) is the internal potential created by the
fluid shell of density ρo and of lower and upper surface bound-
aries rc and rp respectively:
The perturbations of pressure P′(r, θ) obey the relation of the
hydrostatic equilibrium which is of the following form, correct
to first order in P2(cos Θ):
∇P′ = ρo ∇V′ + ρ′o ∇V0 .
Therefore, the Θ-projection of (36) leads to
(30)
P′(r, Θ) = ρo V′(r, Θ) .
(36)
(37)
Finally, since only the variable part of the pressure, i.e. P′, con-
tributes to the normal effective traction f TN
that acts on the mean
3
surface of core, this latter takes the following expression:
f TN
3 (Rc) = P′(rc) = ρo V′(rc) = −ρo Zc R2
c P2 .
(38)
these three forces,
to
The sum of
corresponds
f TN (Rc) = f TN
deforms
mean surface of the core. Using Eqs. (33-34-38), we get:
the
1 (Rc) + f TN
2 (Rc) + f TN
represented on Fig. 4,
traction
the
normal
3 (Rc)
effective
total
that
f TN (Rc) = (ρc Z − ρo Zc) R2
c P2 − ρc gc Rc S 2 P2 ,
(39)
where the expressions of Z and Zc are given by equations (31c)
and (32c) respectively, so that
f TN (Rc) = X P2(cos Θ) ,
where we have denoted by X the following quantity:
X =
2
5
ρc gc Rc 1 −
ρo
ρc!" 5
2
ζc
Rc − S 2 +
3
2
ρo
ρc
(T2 − S 2)# .
(40)
(41)
Fig. 4. Balance of forces that act on the mean surface of the core
r = Rc: f TN
2 (Rc) the loading
of the solid tide and f TN
1 (Rc) are the gravitational forces, f TN
3 (Rc) the hydrostatic pressure.
3.4. Amplitudeoftheverticaldeformation
According to Melchior (1966), a deforming potential U2 of sec-
ond order produces a deformation at each point rc of the surface
of the core which radial component takes the form
ǫrr = (cid:16)8R2
c(cid:17)
c − 3r2
19µ
ρc U2
r2
c
.
(42)
7
Φo(r) = −
r2 T2 P2
3
5
c − r2
2
3
5
+
ρo
ρc
+
gc
3R2
Rc
p − r2
2
− 3R2
Rc −r2 + 3R2
c1 −
gc
ρo
ρc
1
2
r2 S 2 P2!# .
c − V′(r) ,
V(r) = −
where the effective deforming potential is expressed by
+
R2
p
R2
ρo
ρc
Therefore V(r) = U(r) + Φc(r) + Φo(r) has the following ex-
pression, at any point r inside the core:
Z being a constant that depends on the characteristics of the
planet:
V′(r) = −Z r2 P2 ,
+
3
5
ρo
ρc
(T2 − S 2) +
(31a)
(31b)
(31c)
(32a)
We then obtain its expression, correct to first order in S 2P2 or
T2P2, at any point rc = rc er of the surface of the core:
3
5
S 2# .
c + V′(rc) ,
R2
p
R2
ρo
ρc
gc
Rc
Z =
Rc " ζc
V(rc) = −gc Rc 1 −
where
ρo
3
2
3
2
ρc! +
V′(rc) = −Zc R2
Rc " ζc
ρo
ρc
3
5
Rc
gc
+
Zc =
(32b)
Zc being a constant that depends on the characteristics of the
planet:
c P2 ,
(T2 − S 2) −
2
5
S 2# .
(32c)
Chree (1896) showed that the deformation of the core's surface
under the gravitational forces (that derive from the effective de-
forming potential V′) is the same as that which would result
from the outward normal traction f TN
applied at its mean sur-
1
face r = Rc:
f TN
1 (Rc) = ρc Z R2
c P2 .
(33)
3.3. Totaleffectivenormaltractionactingonthesurfaceof
thecore
The mean surface of the core is subjected to two additional
forces induced by both the loading of the core and the loading of
the ocean, tidally deformed.
First, the pressure due to the differential overloading of the de-
formed elastoviscous matter on the mean surface of radius Rc is
given by the product of the local gravity gc, the density of the
core ρc and the solid tidal height RcS 2:
f TN
2 (Rc) = −ρc gc Rc S 2 P2 .
(34)
We also have to take into account the oceanic hydrostatic pres-
sure. Following Zahn (1966) and Remus et al. (2012), we ex-
press all scalar quantities X(r, Θ) as the sum of their spherically
symmetrical profile X0(r) and their perturbation X′(r, Θ) due to
the tidal potential U(r, Θ) ∝ P2(cos Θ):
X(r, Θ) = X0(r) + X′(r, Θ) ≡ X0(r) + X(r) P2(cos Θ) .
(35)
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
Correct to first order in P2, recalling that rc = Rc (1 + S 2P2), this
reduces to:
where
ǫrr =
5
19µ
ρcU2 ,
(43)
Wp being a constant that depends on the planet characteristics:
V′(rp) = −gc Wp P2 ,
(52b)
where ρcU2 is the deforming traction f T,N(Rc) = XP2 applied on
the core. Furthermore, since the amplitude of the displacement
is also given by ǫrr = S 2P2 (Eq. 25a), we have the following
equality
S 2P2 =
XP2 .
(44)
5
19µ
Therefore the relation (27) beetween µ and ¯µ and the expression
(41) of X enable us to relate the deformation of the surfaces of
the core (S 2) and of the ocean (T2) as
S 2 =
1
¯µ 1 −
ρo
ρc!" 5
2
ζc
Rc − S 2 +
3
2
ρo
ρc
(T2 − S 2)# .
(45)
where
By definition, given in Eq. (28), the impedance F is of the form
F =
2
5
(1 + ¯µ)
Rc
ζc
S 2 .
(46)
Since the surface of the planet is an equipotential, the total po-
tential V takes a constant value at any point rp of the surface of
the ocean:
V(rp) ≡ U(rp) + Φ(rp) = cst .
As V − V′ = cst by definition, we get the simplier condition:
(47)
(48)
V′(rp) ≡ U(rp) + Φ′(rp) = cst .
At a point r of the ocean, Φc(r) corresponds to the external po-
tential created by the core:
c 1
Φc(r) = −gc R2
r
+
3
5
R2
c S 2
r3 P2! .
F =
(49)
1 + ¯µ − ρo
ρc
At the same point r, Φo(r) is the internal potential created by the
fluid shell of density ρo and of lower and upper surface bound-
aries rc and rp respectively:
Φo(r) = − gc
R3
p
Rc
+ gc R2
c
3
5
r2T2
R3
p
ρo
ρc
ρc 1
ρo
r
3R2
+
p − r2
2R3
p
R2
c S 2
3
5
+
r3 P2! .
P2
(50)
Therefore V(r) = U(r) + Φc(r) + Φo(r) has the following ex-
pression, at any point r inside the ocean:
V(r) = −gc R2
c 1 −
ρo
ρc! 1
r −
ρo
ρc
r2
2R3
c
+
3
2
ρo
ρc
where the effective deforming potential is expressed by
R2
p
R3
c + V′(r) ,
(51a)
(51b)
V′(r) = −gc R2
c W(r) P2 ,
W being a function of the distance r to the center of the planet:
W(r) =
ζc
Rc
r2
R3
c
+
3
5
ρo
ρc
r2
R3
c
T2 +
r3 S 2 .
(51c)
We then obtain its expression at a point rp = rp er of the surface
of the planet:
c
3
ρo
5 1 −
ρc! R2
Rc + V′(rp) ,
ρo
ρc
R2
p
V(rp) = −gc 1 −
8
ρo
ρc! R2
c
Rp
+
The condition (48) takes then the form:
Wp =
R2
p
R2
c
ζc +
3
ρo
2
5
R2
p
c
R3
ρo
ρc
p S 2
5 1 −
ρc! R4
+−
Rc − 1 −
5 Rc
Rp!5 1 −
ρo Rc
ρc! S 2 +
Rp!3 1 −
5
2
ρo
ρc
3
α = 1 +
ζc
Rc
= −
ρo
ρc! .
c
ρo
Rp T2 .
ρc! R2
(52c)
2
5
ρo
ρc
α T2 ,
(53)
We may eliminate the variable T2 thanks to Eq. (45):
2
5 α
(cid:16)α + 3
+
ρc(cid:17)
2(cid:17)(cid:16)1 − ρo
ρc 1 −
ρc! +
ρo
ρo
3
2
ρo
ρc
3
2 Rc
Rp!5 1 −
2
5
ρo
ρc
α T2 =
×1 + ¯µ −
ρo
ρc!2 .
Injecting this relation into the expression (53) of ζc/Rc, and the
resulting relation in the expression (46) of F, we finally get:
(cid:16)1 − ρo
ρc(cid:17) (1 + ¯µ)(cid:16)1 + 3
2α(cid:17)
Rp(cid:19)5(cid:16)1 − ρo
4α(cid:18) Rc
ρc(cid:17) − 9
ρc(cid:16)1 − ρo
ρc(cid:17)2
+ 3
2
ρo
.
(56)
In the case of a shallow oceanic envelope (Rp ≃ Rc), the height
of the oceanic tide is then given by Rc(T2 − S 2) at the surface of
the core. Using Eqs. (53-45), we obtain the classical expression
of the height of oceanic tide
The height of the solid tidal displacement is given by RcS 2.
Using Eqs. (57-53-45), we obtain its classical expression:
Rc(T2 − S 2) =
ρo
ρc(cid:17) .
¯µζc
5
ρc
1 − ρo
+ ¯µ(cid:16)1 − 3
2 ζc(cid:16)1 − ρo
ρc(cid:17)
+ ¯µ(cid:16)1 − 3
1 − ρo
ρo
ρc
5
5
ρc(cid:17) ,
RcS 2 =
5
2 ζp
1 + ¯µ
RcS 2 =
which reduces to
(54)
(55)
(57)
(58)
(59)
for an ocean-free planet (ρo = 0), which corresponds to that
given by Lord Kelvin (1863). Thus, recalling Eq. (46), we de-
duce that for an oceanless planet, F is unity.
Fig. 5 displays the value of F for three types of planets (i.e.
Earth-, Jupiter- and Saturn-like planets), with a given core (of
fixed size, mass and shear modulus) and a fluid shell of fixed
density but variable depth, so that the size and mass of the whole
planet varies also. The variation of F is represented in function
of Rc/Rp: the smaller this ratio, the higher the ocean depth.
For the Earth, all parameters are well known (see Table 4).
(52a)
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
The values of the ocean density for the Jupiter- and Saturn-like
planets correspond to the ones we may deduce from the well
known values of their global size and mass (see Table 5), and the
much less constrained values of the core size and mass (see mod-
els A of Table 6). Concerning the shear modulus, we used the
value taken as reference by Henning et al. (2009) when studing
the tidal heating of terrestrial exoplanets, i.e. µ = 5 × 1010 Pa.
These models of planets are used as starting points to compare
the influence the ocean depth has on core deformation for dif-
ferent types of planets. Since we do not try to estimate this de-
formation for realistic planets, we will not discuss in this section
the validity of the values we use for the parameters.
Fig. 5 shows that for a planet with a shallow fluid shell
(i.e. when Rc & a × Rp, where a = {0.840, 0.915, 0.937} for an
Earth-, Jupiter- and Saturn-like planet respectively), F is less
than unity, which means that the ocean exerts a loading effect
on the solid core which is stronger than the gravitational forces
and opposed to it. That is the case of the Earth where the depth
of the oceanic envelope does not exceeds 1% of the size of the
planet, but giant planets are supposed to have a solid core no
bigger than the third of the planet size. According to Fig. 5, F
may reach values up to 2.3 for this kind of planets. That means
that for a planet with a deep fluid envelope, the ocean tide has
no loading effect on the core but exerts a gravitational force that
amplifies the tidal deformation. We refer the reader to Dermott
(1979) for a complete discussion.
3.5. ModifiedLovenumbers
From Eq. (56) we deduce the Love number hF
measures the surface deformation:
2 (cf. Eq. 28) which
hF
2 =
1 + ¯µ − ρo
ρc
5
2 (cid:16)1 − ρo
ρc(cid:16)1 − ρo
ρc(cid:17)(cid:16)1 + 3
2α(cid:17)
Rp(cid:19)5(cid:16)1 − ρo
4α(cid:18) Rc
ρc(cid:17) − 9
ρc(cid:17)2
ρo
+ 3
2
.
(60)
Reference
http://nssdc.gsfc.nasa.gov/
planetary/factsheet/earthfact.html
Charette & Smith (2010)
Quantity
Value
5.9736 × 1024
6371 × 103
M⊕p (kg)
R⊕p (m)
M⊕c (kg) M⊕p − 1.4 × 1021
R⊕p − 3682 × 103
R⊕c (m)
Table 4. Earth parameters.
Quantity
Jupiter
Rp (m)
Mp (kg)
10.973 × R⊕p
317.830 × M⊕p
Saturn
9.140 × R⊕p
95.159 × M⊕p
Reference
http://nssdc.gsfc.nasa.
gov/planetary/factsheet/
Table 5. Mass and mean radius of Jupiter and Saturn.
Rc (m)
Mc (kg)
Reference
Jupiter
0.126 × Rp
6.41 × M⊕p
http://www.oca.eu/guillot/
jupsat.html (Guillot, 1998)
Saturn
0.219 × Rp
18.65 × M⊕p
Hubbard et al. (2009)
Table 6. Mass and mean radius of Jupiter's and Saturn's cores.
for Earth-,
in terms of
Factor F, accounting for
Fig. 5.
the overloading ex-
erted by the tidally deformed oceanic envelope on the
solid core,
the ocean depth throw the ratio
Rc/Rp for three types of planet. Parameters are given in
Tables 4-5-6:
Jupiter- and Saturn-like plan-
ets, we
that: Rp = {1, 10.97, 9.14}
(in units of R⊕p), Mp = {1, 317.8, 95.16} (in units of M⊕p ),
(in
Rc = {0.99, 0.126, 0.219}× Rp, Mc = {0.33, 6.41, 18.65}
units of M⊕p ) and µ = 5 × 1010 (Pa) for all cores. Note that this
figure is similar to that drawned by Dermott (1979), divergences
coming from the use of different values of the parameters.
respectively
assume
Let us give here the expression of the second-order Love number
(20) associated with the solid core. First of all, let us recall its
value for an ocean-free planet: According to Eq. (29) with r =
Rc, we get
k2 =
h2 =
3
5
3
2
1
1 + ¯µ
.
(61)
In presence of an ocean on top of the solid core, we may also
introduce the modified Love number
V′(Rc) − U(Rc)
(62)
def
=
=
=
kF
2
V′(Rc)
U(Rc) − 1 ,
Φ′(Rc)
U(Rc)
U(Rc)
where V′(Rc) and U(Rc) are obtained from Eqs. (31b) and (21)
respectively, with r = Rc. Thus, expressing ζc/Rc in function
of the modified second-order Love number hF
2 according to
Eq. (28), and using Eq. (45), we obtain the expression of kF
2 in
terms of hF
2 :
kF
2 =1 +
2
5
¯µ
1 − ρo
ρc hF
2 − 1 .
(63)
In this section we have studied the impact of the presence of a
fluid envelope in the determination of the deformation imposed
on an elastic core under tidal forcing. In the following, we con-
sider that the solid core also presents viscous properties so that
its response to the tidal force exerced by the perturber is no more
immediate, thus inducing dissipation. The next section addresses
the quantification of this conversion of energy, which drives the
dynamical evolution of the whole system.
4. Anelastic tidal dissipation: analytical results
Assuming that the anelasticity is linear, the correspondence prin-
ciple established by Biot (1954) allows us to extend the formu-
lation of the adiabatic elastic problem to the resolution of the
9
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
equivalent dissipative anelastic problem. For initial conditions
taken as zero and similar geometries, the Laplace and Fourier
transforms of the anelastic equations and boundary conditions
are identical to the elastic equations, if the rheological param-
eters and radial functions are defined as complex numbers. We
will then denote
(64)
the complex stress tensor, and
4.2. Caseofatwo-layerplanet
Let us introduce the quantities
A = 1 −
B = 1 −
ρo
ρc! 1 +
ρo
ρc
+
3
2
3
2α! , and:
ρc! −
ρc 1 −
ρo
ρo
9
4α Rc
Rp!5 1 −
≈σσσ ≡ ¯¯σσσ1 + iii ¯¯σσσ2
≈ǫǫǫ ≡ ¯¯ǫǫǫ1 + iii ¯¯ǫǫǫ2
(65)
in the expression (56) of F:
F =
A(1 + ¯µ)
B + ¯µ
,
and let us define its complex equivalent
(75a)
,
(75b)
ρo
ρc!2
(76)
(77)
the complex strain tensor.
The perturbative strain is cyclic, with tidal pulsations σ2,m,p,q.
For sake of clarity, we will use from now on the generic notation
ω ≡ σ2,m,p,q, recalling that there is a large range of tidal frequen-
cies for each term of the expansion of the tidal potential. The
stress and strain tensors take the following form:
≈σσσ(ω) = ( ¯¯σσσ1 + iii ¯¯σσσ2) eeeiiiωt ,
≈ǫǫǫ(ω) = ≈ǫǫǫ0 eeeiiiωt .
The complex rigidity
µ(ω) ≡ µ1(ω) + iii µ2(ω) ,
(66a)
(66b)
(67)
where µ1 represents the energy storage and µ2 the energy loss of
the system, is defined by
µ(ω) ≡
≈σσσ(ω)
≈ǫǫǫ(ω)
.
We may also define the complex effective rigidity
by:
where (see Eq. 27)
µ(ω) ≡ ¯µ1(ω) + iii ¯µ2(ω)
µ(ω) = γ µ(ω) ,
γ =
µ
µ ≡
¯µ
µ
=
19
2ρcgcRc
.
(68)
(69)
(70)
(71)
4.1. Caseofafully-solidplanet
The complex Love number k2 may be expressed in terms of the
complex effective rigidity µ, by:
1
1 + µ(ω)
k2(ω) =
=
3
2
3
2
1
1 + γ(cid:2)µ1(ω) + iiiµ2(ω)(cid:3) ,
in the case of a completely solid planet.
The real part of k2 characterizes the purely elastic deforma-
tion, since −Im (k2) gives the phase lag due to the viscosity.
Therefore, we define the factor of tidal dissipation Q, that repre-
sents the dissipation rate due to viscous friction, by
Then, from equation (72), we deduce that
Q−1(ω) = −Im k2(ω)
(cid:12)(cid:12)(cid:12)k2(ω)(cid:12)(cid:12)(cid:12)
Q(ω) = s1 +"
1
¯µ2(ω)
+
¯µ1(ω)
¯µ2(ω)#2
.
(73)
.
(74)
10
A(1 + µ)
B + µ
.
eF =
The determination of how the presence of an oceanic envelope
will modify the tidal dissipation consists in the determination of
kF
2 , defined as the complex Love number k2 in presence of the
fluid envelope. According to the correspondence principle, this
number is given by the complex Fourier transform of (63), i.e. :
kF
2 (ω) =1 +
From (63) we get then
kF
2 (ω) =
1
(B + ¯µ1)2 + ¯µ2
2
×("(B + ¯µ1) C +
2
5
µ(ω)
1 − ρo
ρc
hF
2 (ω) − 1 .
(78)
3
2α
¯µ1! +
3
2α
¯µ2
2# − iiiA D ¯µ2) ,
(79)
where we made use of the dimensionless quantities α, A and B
previously defined (see respectively Eqs. 54 and 75), C and D
given by:
C =
D =
3
2 1 −
2 1 −
3
is of the form:
Q(ω) = vuut1 +
ρo
ρc!2
,
(80a)
(80b)
(81)
Rp!5 1 −
ρo
ρo
+
9
5
3
ρo
ρc
4α Rc
2α! +
ρc! 1 −
ρc!1 +
Rp!5 .
2α Rc
Q−1(ω) = −Im kF
2 (ω)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)kF
(B + ¯µ1(ω))(cid:16) 2αC
4α2A2D2 1 +
¯µ2(ω)
2 (ω)
9
,
3 + ¯µ1(ω)(cid:17)
2
.
(82)
Thanks to the correspondence principle, one is able to derive this
general expression of the tidal dissipation valid for any rheology.
The obtained formulae explicitly reveal its dependence on the
tidal frequency ω ≡ σ2,m,p,q, as shown, for example, by Remus
et al. (2012), Ogilvie & Lin (2004-2007) for fluid layers.
(72)
Finally, the dissipation factor Q, defined here by
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
4.3. Implementationofananelasticmodel
Since the factor Q depends on the real and imaginary compo-
nents ¯µ1 and ¯µ2 of the complex effective shear modulus µ, we
need to define the rheology of the studied body to express it in
terms of the constitutive parameters of the material.
The anelasticity of a material is evaluated by a quality factor Qa
defined by
Qa(ω) =
.
(83)
µ1(ω)
µ2(ω)
We may express the solid tidal dissipation, given by Q (Eq. 74)
for a fully-solid planet and Q (Eq. 82) in the case of a two-layer
planet, in terms of this factor:
where the time derivative of a given quantity is denoted by a dot.
Recalling the relation (68) this equation becomes:
G ≈ǫǫǫ(ω) µ(ω) + η
d
dt(cid:20) µ(ω) ≈ǫǫǫ(ω)(cid:21) = G η
≈ǫǫǫ(ω) .
(87)
Therefore the real part µ1 and the imaginary part µ2 of the com-
plex shear modulus µ have the following expressions:
µ1(ω) =
µ2(ω) =
η2 G ω2
G2 + η2 ω2 ,
η G2 ω
G2 + η2 ω2 .
(88a)
(88b)
(89)
Q(ω ≡ σ2,m,p,q) = s1 +"
1
¯µ2(ω)
+ Qa(ω)#2
,
(84)
The anelastic quality factor Qa is then given by
Qa(ω) =
µ1(ω)
µ2(ω)
=
η ω
G ≡ ω τM ,
and
Q(ω ≡ σ2,m,p,q) =
s1 +
4α2A2D2"1 + B
9
¯µ2(ω)
+ Qa(ω)! 2
3
αC
¯µ2(ω)
+ Qa(ω)!#2
.
(85)
All previous results are independent of the viscoelastic rheolog-
ical model. We have now to apply these general expressions to
a specific rheology which will depend on the physical properties
of the material. Considering our lack of knowledge on the in-
ternal structure of giant planets, we will implement the simplest
model, namely the Maxwell model, which presents the advan-
tage to involve only two parameters and thus to be easy to use
(Tobie 2003, and Tobie et al. 2005). A critical overview of the
four main rheological models has also been done by Henning et
al. (2009), thus we refer the reader to the three above mentioned
papers for a detailed comparison.
4.4. TheMaxwellmodel.
This model considers a viscoelastic material as a spring-dashpot
serie. The instantaneous elastic response is characterized by a
shear modulus G , and the viscous yielding is represented by a
viscous scalar modulus η (see Fig. 6). Notice that the shear mod-
uli G and µ (introduced in Sect. 2.4) designate the same quan-
tity. We change here the notation to avoid any confusion with the
complex shear modulus µ used to study the anelastic tidal dissi-
pation, and whose real and imaginary parts involve both G and
η.
Fig. 6. Representation of the Maxwell model and its correspond-
ing notations.
The constitutive equation is given by Henning et al. (2009):
G ≈σσσ(ω) + η
≈σσσ(ω) = G η
≈ǫǫǫ(ω) ,
(86)
where τM = η/G is the characteristic time of relaxation of the
Maxwell model. As confirmed by Fig. 7, Eq. (89) shows that Qa
increases linearly with the frequency of the cyclic tidal strain:
the shorter the oscillation period, the lower the dissipation due
to intrinsic viscoelastic properties of the material. Moreover, the
anelastic quality factor is proportional to τM = η/G, so that it
dissipates more if it is more rigid and less viscous.
Fig. 7. Anelastic quality factor Qa of the Maxwell model in
function of the tidal pulsation ω for different values of the viscos-
ity η. G is taken equal to 5 × 1010 Pa (see Henning et. al 2009).
Qa is represented on a logarithmic scale.
Thus, we may express µ2 (Eq. 88b) in terms of the anelastic qual-
ity factor Qa (Eq. 89):
µ2(ω) =
G
(ω τM)−1 + ω τM
.
(90)
In the case of a fully-solid body, we get, from Eqs. (83) and (90),
the imaginary part of the complex Love number k2 (72):
Im hk2(ω)i = −
3γ
2
G ω τM
1 + (ω τM)2 (1 + γG)2
.
(91)
Therefore the dissipation factor Q defined by (74) is of the form:
Q(ω ≡ σ2,m,p,q) =
s1 +( 1
Gγ h(ω τM)−1 + ω τMi + ω τM)2
.
(92)
11
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
In the more general case of a two-layer body, the imaginary part
of the complex Love number kF
2 given by Eq. (79), takes a dif-
ferent form than Eq. (91) because of the presence of the fluid
Q (82) with
envelope, so does the two-layer dissipation factor
respect to its oceanless form (92). To obtain them, one needs to
replace the shear modulus µ and the anelastic quality factor Qa
by their expression in the case of the Maxwell model (Eqs 88
and 89 respectively).
Fig. 8 compares the dissipation of the solid core with and without
the presence of a fluid envelope of variable depth for a Saturn-
like planet, using the parameters given by Tables 5-6-7: as ex-
pected, the difference between the two decreases with the size of
the fluid envelope down to about 0.34 × Rp; but for a thiner fluid
shell, the dissipation get lower than it would be without it.
present knowledge of extrasolar giant planets, and also planets
of our Solar System like Jupiter or Saturn, suffers some uncer-
tainties on the values of these parameters, so that the range of
values taken by core properties of giant planets presents poor
constraints. Moreover, even if the presence of a core in Jupiter is
not yet confirmed (see Guillot 1999-2005), but new data coming
from seismology may provide more constraints on giant plan-
ets internal structure (see Gaulme et al. 2011). Nevertheless, we
will explore the resulting tidal dissipation of such bodies around
values of the structural and rheological parameters taken as ref-
erence and corresponding to those of the literature.
5.1. Baselinestructuralandrheologicalparameters
As reference models, we chose Jupiter and Saturn although their
core parameters are still uncertain. The values of the global sizes
and masses of these planets are those of Table 5.
5.1.1. Size and mass of the core
Presently, two main types of models are available for Jupiter's
interior. The NHKFRB group1 uses a three-layer model with
a thin radiative zone, close to previous models by Saumon
& Guillot (2004), whereas the MHVTB group2 proposes a
new type of Jupiter model that possesses only two layers
(see Militzer et al. 2008). But, as explained in Militzer &
Hubbard (2009), the crucial difference lies in the treatment
of the molecular-to-metallic transition in dense fluid hydro-
gen, that leads to very different conclusions. The first group
predicts a small core of less than 10 M⊕p (Saumon & Guillot
2004), while the second obtains a larger core of 14-18 M⊕p
(Militzer et al. 2008). Among all these models of Jupiter's in-
terior, we choose as reference the adiabatic model with Plasma
Phase Transition (PPT) (http://www.oca.eu/guillot/jupsat.html)3
of Guillot (1998), which is of the first type. It predicts a core
of radius Rc = 0.126 × Rp and mass Mc = 6.41 × M⊕p . Only the
mass of the core of this reference model will be used in what
follows. The core radius will serve just as a first approximation,
as a starting point in our study, since we will present our results
for several core sizes.
inte-
There
rior. According to
the model of Guillot with PPT
(http://www.oca.eu/guillot/jupsat.html)3, Saturn's core may
have a mass of Mc = 6.55 × M⊕p and a size of Rc = 0.174 × Rp.
More recently, Hubbard et al. (2009) infer, from Cassini-
Huygens data,
that Saturn has a larger core in the range
Mc = 15-20× M⊕p and a corresponding radius of more than 20%
of the planet size. We adopt this latter as reference model of
Saturn, with Mc = 18.65 × M⊕p and Rc = 0.219 × Rp.
All these models support core accretion as the standard process
for the formation of giant planets. The corresponding parameters
are listed in Table 6.
different models
Saturn's
also
are
of
this value,
2(cid:12)(cid:12)(cid:12)/ Q and (cid:12)(cid:12)(cid:12)k2(cid:12)(cid:12)(cid:12)/Q. We
Fig. 8. Relative difference between (cid:12)(cid:12)(cid:12)kF
may distinguish two regimes: for 0 < Rc < 0.661 Rp
the presence of the fluid envelope increases the tidal dis-
sipation; over
than
it would have been without fluid shell. Parameters are
given in Tables 5-6-7 for a Saturn-like planet perturbed at
the tidal frequency of Enceladus (ω = 2.25 × 10−4 rad · s−1):
Rp = 9.14 R⊕p, Mp = 95.16 M⊕p, Rc = 0.219 R⊕p, Mc = 18.65 M⊕p ,
G = 5 × 1010 (Pa) and η = 1015 (Pa · s).
tidal dissipation is lower
Planet
Satellite
ω (rad · s−1)
Reference
Jupiter
Io
2.79 × 10−4
Ioannou et Lindzen (1993)
Saturn
Enceladus
2.25 × 10−4
Lainey et al. (2012)
Table 7. Tidal frequencies considered in numerical applications.
5. Anelastic tidal dissipation: role of the structural
and rheological parameters
The main uncertainties concern the viscoelastic properties of
the core, namely its shear modulus G and its viscosity η. At
5.1.2. Rheological parameters of the core
With our choice of the Maxwell model to represent the rheol-
ogy of the solid parts of the planet, the dissipation quality factor
Q depends on the tidal frequency ω and on four structural and
rheological parameters: the relative size of the core (Rc/Rp), the
relative density of the envelope with respect to the core (ρo/ρc),
the shear modulus (G) and the viscosity of the core (η). Our
12
2008).
3 It
1 Nettelmann, Holst, Kietzmann, French, Redmer and Blaschke
(Nettelmann et al. 2008).
2 Militzer, Hubbard, Vorberger, Tamblyn, and Bonev (Militzer et al.
is constructed with CEPAM, Code d'Evolution Plan´etaire
Adaptatif et Modulaire (Guillot & Morel, 1995).
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
Fig. 9. Dissipation quality factor Qeff of the Maxwell model in function of the viscoelastic parameters G and η. Top: for a Jupiter-
like two-layer planet tidally perturbed at the Io's frequency ω ≃ 2.79 × 10−4 rad · s−1. Bottom: for a Saturn-like two-layer planet
Qeff =
tidally perturbed at the Enceladus' frequency ω ≃ 2.25 × 10−4 rad · s−1. The red dashed line corresponds to the value of
{(3.56 ± 0.56) × 104, (1.682 ± 0.540) × 103} (for Jupiter and Saturn respectively) determined by Lainey et al. (2009-2012). The
blue lines corresponds to the lower and upper limits of the reference values taken by the viscoelastic parameters G and η for an
unknown mix of ice and silicates. We assume the values of Rp = {10.97, 9.14} (in units of R⊕p), Mp = {317.8, 95.16} (in units of M⊕p ),
Mc = {6.41, 18.65} × M⊕p given in Tables 5-6.
high pressure and temperature, theoretical models and experi-
ments show that G and η values depend on temperature and pres-
sure. But no experiments are available at the very-high pressures
and temperatures we may expect in Jupiter's and Saturn's cores
(Guillot 2005). Nevertheless, geophysical and experimental data
have allowed to constrain the rheology of the icy satellites of
Jupiter, since their ranges of pressure and temperature are simi-
lar to those of the outer mantle of the Earth (Tobie 2003). Then,
keeping in mind that these values may differ by several orders of
magnitude in our case, we will adopt reference values based on
these data, assuming that Jupiter's and Saturn's cores are made
of ice and rock. We will then explore, in the following Sect. 5.2,
the variation of the tidal dissipation for a large range of values
of the rheological parameters around those taken as reference.
We thus assume that the shear modulus G is in the range
Concerning the viscosity η, it takes its values in the range
hGice = 4 × 109 (Pa) , Gsilicate = 1011 (Pa)i (Henning et al. 2009).
hηice = 1014Pa · s, ηsilicate = 1021 Pa · si for the icy satellites of
Jupiter at high pressure (Tobie 2003). We expand this range, re-
ducing its lower boundary by two orders of magnitude, following
the discussion of Karato (2011) which seems to indicate that, at
the very high pressures, viscosity in deep interior of super-Earths
may decrease by two or three orders of magnitude. We refer
the reader to the Karato's paper for an overview of all plausi-
ble mechanisms that may cause a change in the viscous-pressure
relationship at very-high pressures.
5.2. Dependenceoftidaldissipationonrheology
Since tidal dissipation causes exchange of angular momentum
in the system, it may be quantified by monitoring carefuly the
orbital motion of the system. Using astrometric data covering
more than a century, Lainey et al. (2009-2012) succeeded in
determining from observations the tidal dissipation in Jupiter
and Saturn: namely, QJupiter = (3.56 ± 0.56) × 104 determined by
Lainey et al. (2009), and QSaturn = (1.682 ± 0.540) × 103 deter-
mined by Lainey et al. (2012) and requireded by the formation
scenario of Charnoz et al. (2011). However, with such a method,
the different contributions to the global tidal dissipation, com-
ing from each layer constituting the planet, are lumped together.
Equations of the dynamical evolution (Eqs. 99 to 103) link the
observed evolution rates of the rotational and orbital parame-
ters to the observed tidal dissipation and system characteristics.
Since all these rates are proportional to R5
p, where Rp is the planet
13
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
Fig. 10. Dissipation quality factor Qeff normalized to the size of the planet for Jupiter-like and Saturn-like giant planets. Note that all
curves are represented with a logarithmic scale. Left: dependence to the perturbative strain pulsation ω, with Rc = {0.20, 0.34} × Rp.
Right: dependence to the size of the core, with ω ≃ 2.25 × 10−4 rad · s−1 (tidal frequency of Enceladus) for the blue curve associated
to a Saturn-like planet, and ω ≃ 2.79 × 10−4 rad · s−1 (tidal frequency of Io) for the red curve associated to a Jupiter-like planet. The
green curve corresponds to the prescription of Goodman & Lackner (2009) (see plain text for details). The red and blue dashed
lines correspond to the value of Qeff = {(3.56 ± 0.56) × 104, (1.682 ± 0.540) × 103} (for Jupiter and Saturn respectively) determined
by Lainey et al. (2009-2012). Their zone of uncertainty is also represented in the corresponding color. We assume the values of
Rp = {10.97, 9.14} (in units of R⊕p), Mp = {317.8, 95.16} (in units of M⊕p ), Mc = {6.41, 18.65} × M⊕p given in Tables 5-6. We also take
for the viscoelastic parameters G = {4.85, 4.45} × 1010 (Pa), and η = {1.26, 1.78} × 1014 (Pa · s−1) for Jupiter and Saturn respectively.
radius, we introduce the associated dissipation factor 4:
Qeff = Rp
Rc!5
× kF
2 (Rp)
2 (Rc) × Q ,
kF
(93)
where kF
2 (Rc) can be deduced from Eq. (81), and kF
2 (Rp) des-
ignates the modulus of the second order Love number of the
planet's surface that is obtained from Eqs. (52) and (21)
kF
2 (Rp) =
V′(Rp)
U(Rp) − 1
=
2
5
ρo
ρc
ρo
Rp(cid:19)3 (cid:16)1 − ρo
ρc −(cid:18) Rc
ρc(cid:17)
Rp(cid:19)5(cid:16)1 − ρo
5(cid:18) Rc
ρc(cid:16)α + 3
ρc(cid:17)2 ρo
2(cid:17) 1
α − 3
2 Rc
ρc!2
ρc! +
ρc 1 −
Rp!5 1 −
3
2
ρo
ρo
ρo
3
,
H
where H acounts for the quantity
H = 1 + µ −
ρo
ρc
+
kF
2 (Rc) and kF
the second
Therefore, we can evaluate the real part of
order Love numbers
2 (Rp), accounting re-
spectively for the deformation of the core's and planet's
surface, for parameters whose values are compatible with
the tidal dissipation observations
(Lainey et al., 2009-
in this order, assuming
2012). For Jupiter and Saturn,
G = {4.85, 4.45} × 1010 Pa
Rc = {0.170, 0.260}× Rp,
that
η = {1.26, 1.78} × 1014 Pa · s,
that
and
Re hkF
2 (Rc)i = {3.21, 3.31}
2 (Rp)i = {1.37, 0.24}.
and
These estimations at the planet's surface can be compared to the
value of Gavrilov & Zharkov (1977) of k2 = 0.379 for Jupiter
and k2 = 0.341 for Saturn obtained for stratified models. As
discussed in the aforementioned paper, the differences between
both evaluations are linked to the degree of stratification: the
more the planet interior is stratified, the smaller the second order
Love number (we recall that the second order Love number of a
homogeneous fluid planet is 3/2).
Re hkF
obtain
we
(94)
.
(95)
5.3. Dependenceoftidaldissipationonthesizeofthecore
andthetidalfrequency
Since we have weak constraints on the viscoelastic parameters of
giant planet cores (Guillot 2005), we thus have to explore a large
Qeff
range of values. Fig. 9 shows the tidal dissipation factor
around the reference values presented in Sect. 5.1, expanding
the range up to about ± 2-4 orders of magnitude for G and η. In
the middle region (inside the blue rectangle on Fig. 9), where η
and G correspond to the reference values, the dissipation factor
Qeff of Saturn (resp. Jupiter) may reach values of the order of
103 (resp. 104), and in the whole field it varies up to 1020.
From Fig. 9, we deduce that the tidal dissipation of the core may
reach the values observed for Jupiter (Lainey et al. 2009) and
Saturn (Lainey et al. 2012) assuming that Jupiter's core (resp.
Saturn's core) has a radius 34.92% (resp. 18.72%) larger than
this of Guillot 1998 (resp. Hubbard 2009).
4 This can also be demonstrated by calculating the perturbation of
the gravific potential at the surface of the planet, adapting Zahn (1966)
theory taking into account the boundary conditions at the core surface.
For Fig. 10, the values of the viscoelastic parameters G and η,
and of the core size Rc result from Fig. 9: we chose them so as the
tidal dissipation factor Q reaches the observed values of Lainey
et al. (2009-2012) on the condition that the rheological parame-
ters stand in the more realistic domain defined by the lowest and
highest value of the ice and rock viscoelasticities taken as refer-
ence. Taking into account the global dissipation values obtained
by Lainey et al. 2009 & 2012 for Jupiter and Saturn, we may
deduce some constraints on the viscoelastic parameters and also
the size of the core (looking at the red dashed line in Fig. 9). We
thus will assume that they take values allowing such a dissipa-
tion:
-- we first chose a core slightly larger than that assumed until
now: Rc = 0.170 × Rp for Jupiter, and Rc = 0.260 × Rp for
Saturn,
-- we then fixed the value of the shear modulus G to the low-
est value needed to reach the observed tidal dissipation of
14
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
Lainey et al. (2009-2012), i.e. G = 4.85 × 1010 Pa for Jupiter,
and G = 4.45 × 1010 Pa for Saturn,
-- e finally searched the more realistic value of the viscos-
ity which corresponds to the observed tidal dissipations of
Jupiter and Saturn: η = 1.26 × 1014 Pa · s for Jupiter, and
η = 1.78 × 1014 Pa · s for Saturn.
Fig. 10 explores, for the present model, the dependence of Qeff
on the pulsation ω and the size of the core Rc normalized by the
size of the planet Rp. With these parameters, the figure indicates
that Saturn dissipates slightly more than ten times more than
Jupiter, since (Rc)Saturn > (Rc)Jupiter and (ρc)Saturn < (ρc)Jupiter. In
the range of tidal frequencies of Jupiter's and Saturn's satellites
(2.25 × 10−4 rad · s−1 < ω < 2.95 × 10−4 rad · s−1, Lainey et
Qeff keeps the
al. 2009-2012), the effective dissipation factor
same order of magnitude. However, it strongly depends on the
size of the core, as it may loose up to 6 orders of magnitude
between a coreless planet and a fully-solid one. Note that for
a given core (so that Mc, Rc and then ρc are fixed) and a given
mass of the planet Mp, the density of the fluid envelope ρo, which
varies with its height Rp − Rc, can not exceed ρc. Since:
ρo (Rc) =
c(cid:17) ,
p − R3
this condition gives a limit for the core size:
Mp − Mc
4/3 π(cid:16)R3
= Mc
Mp!1/3
Rp!sup
Rc
(96)
.
(97)
In 2004, Ogilvie & Lin also studied tidal dissipation in giant
planets, and particulary the tidal dissipation resulting from the
excitation of inertial waves in the convective region by the tidal
potential for rotating giant planets with an elastic solid core.
They obtained a decrease of the quality factor Q of one or-
der of magnitude considering the dynamical tide (due to iner-
tial modes) compared to the equilibrium one: from Q = 106 to
Q = 105, but it is not efficient enough to explain the observed
tidal dissipations in Jupiter or Saturn which are of 1-2 orders
of magnitude higher (Lainey et al. 2009-2012). Moreover, they
showed that the dissipation resulting from the resonance be-
tween fluid tide and inertial modes highly depends on the tidal
frequency in the range of inertial waves, as do also the core-
less models (Wu 2005). This disagrees with the weak frequency-
dependence obtained from astrometry (Lainey et al. 2012).
By discussing the size of the core, Goodman & Lackner (2009)
got a higher value of the quality factor Q, in the range 107 −
108 ×(cid:16)0.2 Rp/Rc(cid:17)5
which disagrees with the observed value of
the tidal dissipation of Saturn (see Fig. 9).
The present two-layer model proposes an alternative process
to reach such a high dissipation with a smooth frequency-
dependence of Q. But the uncertainties on the structural and rhe-
ological parameters does not allow us to firmly conclude that the
tidal dissipation of the core can explain by its own the tidal dis-
sipation observed in giant planets of our Solar System by Lainey
et al. (2009-2012).
Through these expressions of the tidal dissipation closely linked
to the internal structure of the planet and its rheological proper-
ties, we are now able to derive the equations of the dynamical
evolution of the system with explicit dependence on the tidal
frequency.
5.4. ComparisonwithpreviousworkofDermott
The difference beetwen our study and the work of Dermott
(1979) lies in the treatment of the tidal dissipation.
Dermott draws his expression from an evolution scenario of
Saturnian and Jovian systems. He assumes that the satellites of
Jupiter and Saturn were formed 4.5 × 109 ago, and that their
semi-major axis have changed by 10 % since their formation,
with a stable resonance of the main satellites of Jupiter (Io,
Europa and Ganymede) since their formation but a young res-
onance of the satellites Mimas and Thetys of Saturn. He also
assumes an average value of the tidal dissipation, independent
of amplitude, frequency and time. All these assumptions lead
Dermott (1979) to a tidal dissipation factor Q that only depends
on the mass Mc, the size Rc, the elasticity µ of the core and a
dimensionless coefficient K characteristic of the evolution sce-
nario of the planet. In particular, his tidal factor Q is directly
proportional to R5
c (Eq. 27 of Dermott 1979), so that the tidal
dissipation gets lower as the core size increases (see Fig. 4 of
Dermott 1979), while one should expect an opposite behaviour.
Instead, our model is constructed on physical considerations of
the internal structure and properties of the core. In particular, we
derived our tidal dissipation factor Q with no assumption on the
evolution of the Jovian and Saturnian systems. To do so, we used
the correspondence principle of Biot (1954). This allowed us to
obtain an expression of the tidal dissipation factor valid for any
rheological model of planets' cores. Moreover, our expression
(82) of Q not only depends on the mass Mc, the size Rc and the
elasticity µ of the core, but also on the tidal frequency ω and
the viscosity η. We notice, in particular, that tidal dissipation in-
creases when the size of the core increases, contrary to Dermott's
strange result.
6. Equations of the dynamical evolution
Mass redistribution due to the tide generates a tidal torque of
non-zero average which induces an exchange of angular momen-
tum between the orbital motion and the rotation of each com-
ponent. As shown in MLP09 & Remus et al. (2012), this tidal
2(cid:12)(cid:12)(cid:12)/Q (see
torque is proportional to the tidal dissipation ratio(cid:12)(cid:12)(cid:12)kF
ducing a torque,ZV
also Correia & Laskar 2003, Correia et al. 2003, and Murray
& Dermott 2000). Note that for a perfectly elastic material, the
core will be elongated in the direction of the line of centers, in-
r ∧ (ρc∇∇∇U) dV, with periodic variations of
zero average, so that no secular exchanges of angular momentum
will be possible (see Zahn 1966a and Remus et al. 2012). On the
other hand, if the core is anelastic, the deformation of the core
resulting from the equilibrium adjustment presents a time delay
∆t with respect to the tidal forcing, which may be measured also
by the tidal lag angle 2δl or equivalently by the quality factor Q
(see Ferraz-Mello et al. 2008 or Efroimsky & Williams 2009):
tanh∆t × σ2,m,p,qi = tanh2δ(cid:16)σ2,m,p,q(cid:17)i =
1
Q(cid:16)σ2,m,p,q(cid:17) .
(98)
Thus, the tidal bulge is no more aligned with the line of centers,
as shown in Fig. 11.
The resulting tidal angle is at the origin of a torque of non-zero
average which causes exchange of spin and orbital angular mo-
mentum between the components of the system.
The evolution of the semi-major axis a, of the eccentricity e, of
the inclination I, of the obliquity ε and of the angular veloc-
ity Ω ( ¯IA denotes the moment of inertia of A), is governed by
15
F. Remus, S. Mathis, J.-P. Zahn & V. Lainey: Anelastic tidal dissipation in multi-layer planets
where the functions Hm, j,p,q(e, I, ε) are expressed in terms of
d2
j,m(ε), F2, j,p(I), and G2,p,q(e), which are defined in Sect. 2.2
H2,m, j,p,q(e, I, ε) = s 5
4π
j,m(ε) F2, j,p(I) G2,p,q(e) ,
(104)
(2 − j)!
(2 + j)!
× d2
and Req designates the equatorial radius of body A.
From these equations one may derive the characteristic times of
synchronization, circularization and spin alignment:
1
tsync
1
tcirc
1
talignA
1
talignOrb
= −
,
1
= −
Ω − n
de
1
dt
e
1
ε
1
I
dε
dt
dI
dt
= −
= −
dΩ
dt
= −
1
¯IA (Ω − n)
d(cid:16) ¯IA Ω(cid:17)
dt
,
=
=
1
d (cos ε)
ε sin ε
dt
1
d (cos I)
I sin I
dt
,
.
(105)
(106)
(107)
(108)
7. Conclusion
2 and kF
To conclude, the purpose of this work was to study the tidal dis-
sipation in a two-layer planet consisting in a rocky/icy core and
a fluid envelope, as one expects to be the case in Jupiter, Saturn,
and many extrasolar planets. We considered the most general
configuration where the perturber (star or satellite) is moving on
an elliptical and inclined orbit around the planet which rotates on
an inclined axis. We expanded the tidal displacement in Fourier
series and spherical harmonics, each term of the expansion hav-
ing a radial part proportional to that of the corresponding term
of the tidal potential, which depends on the eccentricity, the in-
clination and the obliquity. We followed the method by Dermott
(1979) to derive the modified Love numbers hF
2 account-
ing for the tidal deformation at the boundary of the solid core.
Like Dermott, we made the simplifying assumption that core and
envelope have a constant density. Then, generalizing the results
of his work invoking the correspondence principle, we obtained
2 / Q, Q be-
the tidal dissipation rate of the core expressed by kF
ing the quality factor. That ratio depends on the tidal frequency
and on the rheological properties of the core; unlike Dermott we
made no assumption on the formation history of the system. As
mentioned in Sect. 5.1 the rheological properties of planetary
cores are still quite uncertain. However, taking plausible values
for the viscoelastic parameters G and η, we obtain a tidal dissi-
pation which may be much higher than for a fully fluid planet
and weakly frequency-dependent. Under these assumptions, we
find that the low value of Q = (1.682 ± 0.540) × 103, determined
by Lainey et al. (2012) and needed by Charnoz et al. (2011) to
explain the formation of all mid-size moons of Saturn from the
rings, can be reached taking into account the tidal dissipation of
Saturn's core. In the same way, the dissipation in Jupiter's core
may explain the value of the Q-factor determined by Lainey et
al. (2009), i.e. Q = (3.56 ± 0.56) × 104. But to do so, we need to
assume a core in Jupiter and Saturn slightly larger than those of
Guillot (1998) and Hubbard et al. (2009). But we recall that in
our model the density was assumed to be piecewise constant. In
the future, we shall consider a non constant density profile, to
evaluate the impact on our results of a realistic density stratifi-
cation. Moreover, there are much uncertainties on the determi-
nation of the core size in giant planets as Jupiter and Saturn, so
Fig. 11. Tidal interaction involving a solid body. Body B exerts
a tidal force on body A, which adjusts itself with a phase lag
2 δ, because of internal friction in the anelastic core. This ad-
justment may be split in an adiabatic component, corresponding
to the elastic deformation, in phase with the tide, and a dissipa-
tive one, resulting from the viscous internal frictions, which is in
quadrature.
the following equations, established in MLP09 and Remus et al.
(2012):
2 (Rp, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
(cid:12)(cid:12)(cid:12)kF
Qeff(σ2,m,p,q)
BR5
GM2
eq
a6
2 (Rp, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
Qeff(σ2,m,p,q)
=
dt
dt
¯IA Ω
4π
5
8π
5
= −
d (cos ε)
BR5
GM2
eq
a6
d(cid:16) ¯IA Ω(cid:17)
× X(m, j,p,q)∈I
( j + 2 cos ε)(cid:12)(cid:12)(cid:12)kF
× X(m, j,p,q)∈I
× X(m, j,p,q)∈I
(2 − 2p + q)(cid:12)(cid:12)(cid:12)kF
GMBR5
a8
= −
4π
5
da
dt
1
a
2
n
eq
2 (Rp, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
Qeff(σ2,m,p,q)
(99)
,
hH2,m, j,p,q(e, I, ε)i2
hH2,m, j,p,q(e, I, ε)i2
hH2,m, j,p,q(e, I, ε)i2
,
,
(100)
(101)
1
e
de
dt
= −
d (cos I)
dt
16
eq
1
1
n
4π
5
1 − e2
e2
Qeff(σ2,m,p,q)
GMBR5
a8
× X(m, j,p,q)∈I
"(2 − 2p) 1 −
× (cid:12)(cid:12)(cid:12)kF
2 (Rp, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
√1 − e2
√1 − e2! + q#
hH2,m, j,p,q(e, I, ε)i2
× X(m, j,p,q)∈I(cid:2) j + (2q − 2) cos I(cid:3)
× (cid:12)(cid:12)(cid:12)kF
2 (Rp, σ2,m,p,q)(cid:12)(cid:12)(cid:12)
hH2,m, j,p,q(e, I, ε)i2
BR5
GM2
eq
a8
Qeff(σ2,m,p,q)
4π
5
1
n
1
=
,
(102)
,
(103)
that we need more constraints on the system formation by core
accretion (Pollack et al. 1996) and differenciation resulting from
the internal structure evolution (Nettelmann 2011). Furthermore,
seismology seems to offer an interesting way to improve our
knowledge of giant planets interiors (Gaulme et al. 2011).
To conclude, the purpose of this paper was to study tidal dissi-
pation in the solid parts of a simple model of two-layer planet;
we show how this mechanism may be powerful. The results de-
rived here are general in the sense that no specific rheological
model has been assumed. However, considering the lack of con-
straints on the rheology of giant planets cores, we have chosen
the simplest Maxwell model to illustrate the tidal dissipation.
This work represents thus a first step for further numerical inves-
tigations in more realistic cases.
Acknowledgments
The authors are grateful to the referee for his/her remarks and
suggestions. They also thank G. Tobie for fruitful discussions
during this work and T. Guillot for providing numerical models
of Jupiter and Saturn interiors. This work was supported in part
by the Programme National de Plan´etologie (CNRS/INSU), the
EMERGENCE-UPMC project EME0911, and the CNRS pro-
gramme Physique th´eorique et ses interfaces.
References
Alterman, Z., Jarosch, H., & Pekeris, C. L. 1959, Royal Society of London
Proceedings Series A, 252, 80
Biot, M. A. 1954, Journal of Applied Physics, 25, 1385
Charnoz, S., Crida, A., Castillo-Rogez, J. C., et al. 2011, Icarus, 216, 535
Charette, M.A., & Smith, W.H.F. 2010, Oceanography, 104, 106
Chree, C. 1896, Cambridge Phil. Trans., 16, 14
Correia, A. C. M., Laskar, J., & de Surgy, O. N. 2003, Icarus, 163, 1
Correia, A. C. M., & Laskar, J. 2003, Icarus, 163, 24
Dahlen, F. A., Tromp, J., & Lay, T. 1999, Physics Today, 52, 61
Dermott, S. F. 1979, Icarus, 37, 310
Efroimsky, M., & Williams, J. G. 2009, Celestial Mechanics and Dynamical
Astronomy, 104, 257
Ferraz-Mello, S., Rodriguez, A., & Hussmann, H. 2008, Celestial Mechanics and
Dynamical Astronomy, 101, 171
Gaulme, P., Schmider, F.-X., Gay, J., Guillot, T., & Jacob, C. 2011, A&A, 531,
A104
Gavrilov, S. V., & Zharkov, V. N. 1977, Icarus, 32, 443
Goodman, J., & Lackner, C. 2009, ApJ, 696, 2054
Greff-Lefftz, M., M´etivier, L., & Legros, H. 2005, Celestial Mechanics and
Dynamical Astronomy, 93, 113
Guillot, T., Chabrier, G., Gautier, D., & Morel, P. 1995, ApJ, 450, 463
Guillot, T., & Morel, P. 1995, A&AS, 109, 109
Guillot, T. 1999, Planet. Space Sci., 47, 1183
Guillot, T. 2005, Annual Review of Earth and Planetary Sciences, 33, 493
Henning, W. G., O'Connell, R. J., & Sasselov, D. D. 2009, ApJ, 707, 1000
Hubbard, W. B., Dougherty, M. K., Gautier, D., & Jacobson, R. 2009, Saturn
from Cassini-Huygens, 75
Karato, S.-I. 2011, Icarus, 212, 14
Kaula, W. M. 1962, AJ, 67, 300
Lainey, V., Arlot, J.-E., Karatekin, O., & van Hoolst, T. 2009, Nature, 459, 957
Lainey, V., Karatekin, O., Desmars, J., Charnoz, S., Arlot, J.-E., Emelyanov, N.,
Le Poncin-Lafitte, C., Mathis, S., Remus, F., Tobie, G., & Zahn, J.-P. 2012,
submited to ApJ
Lambeck, K. 1980, The Earth's Variable Rotation (Cambridge University Press)
Love, A. E. H. 1911, Some Problems of Geodynamics (Cambridge University
Press)
Mathis, S., & Le Poncin-Lafitte, C. 2009, A&A, 497, 889
Mathis, S., & Zahn, J.-P. 2005, A&A, 440, 653
Melchior, P. 1966, The Earth Tides (Pergamon, New York)
Militzer, B., Hubbard, W. B., Vorberger, J., Tamblyn, I., & Bonev, S. A. 2008,
ApJ, 688, L45
Militzer, B., & Hubbard, W. B. 2009, Ap&SS, 322, 129
Murray, C. D., & Dermott, S. F. 2000, Solar System Dynamics (Cambridge
University Press)
Nettelmann, N., Holst, B., Kietzmann, A., et al. 2008, ApJ, 683, 1217
Nettelmann, N. 2011, Ap&SS, 336, 47
Ogilvie, G. I., & Lin, D. N. C. 2004, ApJ, 610, 477
Ogilvie, G. I., & Lin, D. N. C. 2007, ApJ, 661, 1180
Ogilvie, G. I. 2009, MNRAS, 396, 794
Peale, S.J., & Cassen, P. 1978, Icarus, 36, 245
Pollack, J. B., Hubickyj, O., Bodenheimer, P., et al. 1996, Icarus, 124, 62
Remus, F., Mathis, S. & Zahn, J.-P. 2012, submited to A&A
Rieutord, M. 1987, Geophysical and Astrophysical Fluid Dynamics, 39, 163
Ross, M., & Schubert, G. 1986, Lunar and Planetary Science Conference
Proceedings, 16, 447
Santos, N. C., & et al. 2007, JENAM-2007, "Our Non-Stable Universe"
Saumon, D., & Guillot, T. 2004, ApJ, 609, 1170
Takeuchi, H., and M. Saito 1972, Seismic surface waves, Methods Comput.
Phys., 11, 217295
Thomson, W. (Lord Kelvin) 1863, Dynamical problems regarding elastic
spheroidal shells, and On the rigidity of the Earth, Phil. Trans. Roy. Soc.
London 153, 573-616.
Tobie, G. 2003, Impact du chauffage de mar´ee sur l'´evolution g´eodynamique
d'Europe et de Titan, PhD thesis, Universit´e Paris 7 - Denis Diderot
Tobie, G., Mocquet, A., & Sotin, C. 2005, Icarus, 177, 534
Wu, Y. 2005, ApJ, 635, 688
Yoder, C. F. 1995, Icarus, 117, 250
Zahn, J.-P. 1966a, Annales d'Astrophysique, 29, 313
Zahn, J. P. 1966b, Annales d'Astrophysique, 29, 489
Zahn, J.-P. 1977, A&A, 57, 383
Appendix: elastic system
In Sect. 2.3 we gave the system of equations (11), with its bound-
ary conditions (12), governing an elastic planet under tidal per-
turbation. Considering that such perturbations are of small order
of magnitude compared to the hydrostatic equilibrium, we pro-
posed in Sect. 2.4 a method to linearize the system (14). Thus,
assuming the expansion (18) of all quantities in spherical har-
monics, we obtain the following system of equations governing
the scalar radial parts of these expansions (Alterman et al. 1959;
Takeuchi & Saito 1972):
ym
2 +
3 µ
ym
3
r
3 µ
4µK
K + 4
6(cid:16)K − 2
3 µ(cid:17)
(cid:16)K + 4
3 µ(cid:17) r
ym
4
r − ρ0ym
6 ,
ym
2
r
ym
1
r −
ym
3
r
+ 6
ym
ym
6µK
12µK
+
ym
1
r
+
K + 4
3 µ
r
R
ym
1 = −
1
K + 4
r
R −
ym
3
r
2(cid:16)K − 2
3 µ(cid:17)
3 µ(cid:17) r
2 =−4 ρ0 gs
(cid:16)K + 4
3 µ(cid:17) r
+ 6ρ0 gs
(cid:16)K + 4
4 =ρ0 gs
3 µ(cid:17) r
(cid:16)K + 4
2µh11(cid:16)K − 2
3 µ(cid:17) + 10 µi
(cid:16)K + 4
3 µ(cid:17) r
ym
5
r − 2
5 =4πGρ0ym
ym
ym
6 = − 24πGρ0
1 + ym
6 ,
ym
3
r
ym
3 = −
r
R −
ym
4
µ
,
ym
1
r
+
6
r
ym
1
r
+
+
6µK
+
ym
3
r
,
ym
6
r
.
Solutions of (109) are given by (19).
, (109a)
(109b)
(109c)
(109d)
(109e)
(109f)
17
|
1307.7517 | 1 | 1307 | 2013-07-29T09:47:53 | Physical Properties of Asteroid (308635) 2005 YU55 derived from multi-instrument infrared observations during a very close Earth-Approach | [
"astro-ph.EP"
] | The near-Earth asteroid (308635) 2005 YU55 is a potentially hazardous asteroid which was discovered in 2005 and passed Earth on November 8th 2011 at 0.85 lunar distances. This was the closest known approach by an asteroid of several hundred metre diameter since 1976 when a similar size object passed at 0.5 lunar distances. We observed 2005 YU55 from ground with a recently developed mid-IR camera (miniTAO/MAX38) in N- and Q-band and with the Submillimeter Array (SMA) at 1.3 mm. In addition, we obtained space observations with Herschel/PACS at 70, 100, and 160 micron. Our thermal measurements cover a wide range of wavelengths from 8.9 micron to 1.3 mm and were taken after opposition at phase angles between -97 deg and -18 deg. We performed a radiometric analysis via a thermophysical model and combined our derived properties with results from radar, adaptive optics, lightcurve observations, speckle and auxiliary thermal data. We find that (308635) 2005 YU55 has an almost spherical shape with an effective diameter of 300 to 312 m and a geometric albedo pV of 0.055 to 0.075. Its spin-axis is oriented towards celestial directions (lam_ecl, beta_ecl) = (60 deg +/- 30deg, -60 deg +/- 15 deg), which means it has a retrograde sense of rotation. The analysis of all available data combined revealed a discrepancy with the radar-derived size. Our radiometric analysis of the thermal data together with the problem to find a unique rotation period might be connected to a non-principal axis rotation. A low to intermediate level of surface roughness (r.m.s. of surface slopes in the range 0.1 - 0.3) is required to explain the available thermal measurements. We found a thermal inertia in the range 350-800 Jm^-2s^-0.5K^-1, very similar to the rubble-pile asteroid (25143) Itokawa and indicating a mixture of low conductivity fine regolith with larger rocks and boulders of high thermal inertia on the surface. | astro-ph.EP | astro-ph | Astronomy&Astrophysicsmanuscript no. nea2005YU55astroph
June 30, 2018
c(cid:13) ESO 2018
Physical Properties of Asteroid (308635) 2005 YU55 derived from
multi-instrument infrared observations during a very close
Earth-Approach
T. G. Muller1, T. Miyata2, C. Kiss3 M. A. Gurwell4 S. Hasegawa5, E. Vilenius1, S. Sako2, T. Kamizuka2, T.
Nakamura6, K. Asano2, M. Uchiyama2, M. Konishi2, M. Yoneda7, T. Ootsubo8, F. Usui5, Y. Yoshii2, M. Kidger9, B.
Altieri9, R. Lorente9, A. P´al3, L. O'Rourke9, and L. Metcalfe9
1 Max-Planck-Institut fur extraterrestrische Physik, Giessenbachstrasse, Postfach 1312, 85741 Garching, Germany
2 Institute of Astronomy, School of Science, the University of Tokyo, 2-21-1 Osawa, Mitaka, Tokyo 181-0015, Japan e-mail:
[email protected]
H-1121 Budapest, Hungary
3 Konkoly Observatory, Research Center for Astronomy and Earth Sciences, Hungarian Academy of Sciences; Konkoly Thege 15-17,
4 Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA
5 Institute of Space and Astronautical Science, Japan Aerospace Exploration Agency, 3-1-1 Yoshinodai, Sagamihara, Kanagawa
6 Department of Astronomy, Graduate School of Science, The University of Tokyo, Hongo 7-3-1, Bunkyo-ku, Tokyo 113-0033,
7 Planetary Plasma and Atmospheric Research Center, Tohoku University, Aramaki, Aoba-ku, Sendai 980-8578, Japan
8 Astronomical Institute, Graduate School of Science, Tohoku University, Aramaki, Aoba-ku, Sendai 980-8578, Japan
9 European Space Astronomy Centre (ESAC), European Space Agency, Apartado de Correos 78, 28691 Villanueva de la Canada,
229-8510, Japan
Japan
Madrid, Spain
Received ; accepted
ABSTRACT
The near-Earth asteroid (308635) 2005 YU55 is a potentially hazardous asteroid which was discovered in 2005 and passed Earth on
November 8th 2011 at 0.85 lunar distances. This was the closest known approach by an asteroid of several hundred metre diameter
since 1976 when a similar size object passed at 0.5 lunar distances. We observed 2005 YU55 from ground with a recently developed
mid-IR camera (miniTAO/MAX38) in N- and Q-band and with the Submillimeter Array (SMA) at 1.3 mm. In addition, we obtained
space observations with Herschel⋆/PACS at 70, 100, and 160 µm. Our thermal measurements cover a wide range of wavelengths from
8.9 µm to 1.3 mm and were taken after opposition at phase angles between -97◦ and -18◦. We performed a radiometric analysis via a
thermophysical model and combined our derived properties with results from radar, adaptive optics, lightcurve observations, speckle
and auxiliary thermal data. We find that (308635) 2005 YU55 has an almost spherical shape with an effective diameter of 300 to 312 m
and a geometric albedo pV of 0.055 to 0.075. Its spin-axis is oriented towards celestial directions (λecl, βecl) = (60◦ ± 30◦, -60◦ ±
15◦), which means it has a retrograde sense of rotation. The analysis of all available data combined revealed a discrepancy with the
radar-derived size. Our radiometric analysis of the thermal data together with the problem to find a unique rotation period might be
connected to a non-principal axis rotation. A low to intermediate level of surface roughness (r.m.s. of surface slopes in the range 0.1
- 0.3) is required to explain the available thermal measurements. We found a thermal inertia in the range 350-800 Jm−2s−0.5K−1, very
similar to the rubble-pile asteroid (25143) Itokawa and indicating a mixture of low conductivity fine regolith with larger rocks and
boulders of high thermal inertia on the surface.
Key words. Minor planets, asteroids: individual -- Radiation mechanisms: Thermal -- Techniques: photometric -- Infrared: planetary
systems
1. Introduction
The Apollo- and C-type asteroid (308635) 2005 YU55 is on
a Mars-Earth-Venus crossing orbit1 (Vodniza & Pereira 2010;
Hicks et al. 2010; Somers et al. 2010). Arecibo radar measure-
ments in April 2010 have shown that 2005 YU55 is a very dark,
⋆ Herschel is an ESA space observatory with science instruments
provided by European-led Principal Investigator consortia and with im-
portant participation from NASA.
1 2005, M.P.E.C. 2005-Y47:
nearly spherical object2. They estimated a diameter of about
400 m, in contradiction to earlier calculations based on the V-
magnitude in combination with a low albedo which led to a di-
ameter of only 250 m.
2005 YU55 had a very close Earth approach in November
2011 when it passed within 0.85 lunar distances (0.85 LD) of
the Earth. Later, in January 2029, the asteroid will pass about
0.0023 AU (equivalent to 0.89 LD) from Venus. This close en-
counter with Venus will determine how close the object will pass
2 NASA Near Earth Object Program News:
http://www.minorplanetcenter.org/mpec/K05/K05Y47.html
http://neo.jpl.nasa.gov/news/news171.html
1
3
1
0
2
l
u
J
9
2
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
7
1
5
7
.
7
0
3
1
:
v
i
X
r
a
Muller et al.: MIR Observations of (308635) 2005 YU55
the Earth in 2041 and 20453. The JPL Horizons system gives the
absolute magnitude of 2005 YU55 as H=21.1 mag4. No other as-
teroid with H<23 mag has been observed before to pass inside
1 LD. According to recent orbit simulations it does not pose any
risk of an impact with Earth for the next 100 years5. The closest
recorded approach by an asteroid of similar characteristics was
that of 2004 XP14 (H=19.4 mag) to 1.1 LD on 2006 July 3, hence
the encounter with 2005 YU55 was an exceptional event.
The close Earth approach in November 2011 offered a sev-
eral day observing opportunity from ground and also a brief
(∼16 h) observing window for the Herschel Space Observatory
located in the Lagrangian point L2 at about 1.5 Mio km from
Earth. This was a unique opportunity to study a potentially haz-
ardous asteroid (PHA) in great detail to derive physical and
thermal properties which are needed to make long-term orbit
predictions and to improve our knowledge on Apollo asteroids
in general. We observed this near-Earth asteroid from ground
at mid-Infrared N- and Q-band (miniTAO/MAX38 camera), at
millimetre wavelength (Smithsonian Astrophysical Observatory
Submillimeter Array, or SMA) and from space with Herschel-
PACS at far-infrared wavelengths. We present our observations
(Section 2), the thermophysical model (TPM) analysis (Section
3) and discuss the results (Section 4). In this work, we also con-
sidered a set of auxiliary data (radar, optical, UV and thermal
measurements) which were only available via unrefereed ab-
stracts, astronomical circulars and telegrams.
of the detector array is used for spectroscopy). The MAX38 ob-
serving periods (2011-Nov-08 23:04 to Nov-09 01:51 UT and
from Nov-09 23:56 to Nov-10 02:04 UT) covered the time of the
closest approach (2011-Nov-08 23:24 UT) and about 24 hours
later. The weather conditions were excellent through the ob-
servations. Imaging observations in the 8.9 µm (∆λ ∼ 0.9 µm),
12.2 µm (0.5 µm), and 18.7 µm (0.9 µm) bands were carried out.
α Tuc (IRAS22150-6030, HR 8502, HD 211416) was also ob-
served after the observations of the asteroid as a flux standard.
The absolute flux value of the standard star was obtained via the
α Tuc template spectrum (Cohen et al. 1999).
Since the distance to the asteroid from the Earth was very
short, the asteroid had a very high apparent motion on the sky.
We pointed the telescope at repeated intervals to follow the as-
teroid's movement. The intervals were set to 1 minute and 3 min-
utes on Nov. 8 and 9, respectively. Normal sidereal tracking was
applied in the period between the telescope pointings. Images
were taken at a frame rate of 3.8 Hz with an effective integration
time of 0.197 sec. The frame rate is fast enough not to extend
the image of the asteroid on each frame. Chopping technique6
was not applied because background can be canceled out with
using frames just before or after an object frame. The observa-
tion parameters are summarized in Table 1 and three examples
of reduced images taken at the time of the closest approach are
shown in Fig. 1.
2. Observations
2.1. Groundbasedmid-IRobservationswithMAX38
Table 1. MAX38 Observation Parameters. The first period in-
cludes the closest Earth approach.
2011-Nov-.. (UT)
Day Start
End
Filter
Band
# of
frames
airmass-range
2005 YU55
α Tuc
08 23:04
09 00:30
09 01:11
09 23:56
10 01:38
00:15
00:41
01:24
00:50
02:04
18.7
8.9
12.2
18.7
8.9
20196
2550
1734
12288
4928
1.31-1.47
1.51-1.55
1.65-1.70
2.00-1.61
1.43-1.38
1.41-1.48
1.29
1.31
1.53-1.64
1.67
On Nov. 8, the asteroid was so bright (observatory-centric
distance was about 0.0021-0.0023AU) that it was detectable in
each frame. Sky frames -composed by averaging 7 frames taken
within 2 sec- were subtracted. This successfully canceled out sky
variation similar to observations taken in chopping technique.
Aperture photometry with an aperture radius of 3 arcsec was
applied for each frame. Photometric values were determined by
averaging frames taken in a period of 5 minutes. Errors were
estimated as standard deviation of the photometric values.
On Nov. 9 the asteroid was already at about 0.0084-
0.0089 AU distance and it was difficult to detect the asteroid
on each frame. Here, we added 92 frames into one image by
shifting the frames to compensate the asteroid's movement on
the sky and subtracted the averaged sky frames. The frame-to-
frame shifts were calculated from the ephemeris provided by the
NASA Horizons web page7. The asteroid images in the co-added
frames appeared nearly point-like and no noticeable extensions
6 the telescope's secondary mirror is oscillated between two positions
on the sky at a frequency of a few Hz.
7 JPL Horizons: http://ssd.jpl.nasa.gov/horizons.cgi
Fig. 1. Mid-infrared images of 2005 YU55 obtained by the mini-
TAO/MAX38 camera in the 18.7 µm filter during the time of the
closest Earth approach. North is up and west is right. The aster-
oid moved from right (west) to left (east).
We observed the asteroid 2005 YU55 in the period Nov. 8 -
10, 2011 with the mid-infrared camera MAX38 (Miyata et al.
2008; Nakamura et al. 2010; Asano et al. 2012) attached on the
mini-TAO 1 meter telescope (Sako et al. 2008) which is located
at 5640 m altitude on the summit of Co. Chajnantor in Chile,
which is part of the University of Tokyo Atacama Observatory
Project (PI: Yuzuru Yoshii; Yoshii et al. 2010). MAX38 has a
128×128 Si:Sb BIB detector with a pixel scale of 1.26 arcsec
and a field of view of 2×2.5 arcmin determined by the rectan-
gular field stop in the cold optics (the remaining 0.5×2.5 arcmin
3 http://echo.jpl.nasa.gov/asteroids/2005YU55/2005YU55 planning.html
4 JPL Horizons: http://ssd.jpl.nasa.gov/horizons.cgi
5 JPL's NEO Radar Detection Program Webpage:
http://echo.jpl.nasa.gov/asteroids/index.html
2
Muller et al.: MIR Observations of (308635) 2005 YU55
were detected. We applied aperture photometry on the final sky-
subtracted images. The final flux and error values were again
obtained by averaging all individual photometric values.
In addition to the photometric error we also added a 5% ab-
solute flux calibration error for the N-band data and a 7% error
for the Q-band data based on the radiometric tolerance discus-
sion in Cohen et al. (1999) and the information given in the stel-
lar template. These errors also include possible colour-terms (es-
timated to be below 2%) due to the different spectral shapes of
the star and the asteroid in the N- and Q-band filters. In the first
night (8/9 Nov.) α Tuc and the asteroid were observed at simi-
lar airmass and similar PWV8-levels (based on APEX measure-
ments) and no additional corrections were needed. In the second
night (9/10 Nov.) 2005 YU55 was observed in Q-band at a large
airmass close to 2.0 and a PWV of around 0.5 mm, while α Tuc
was taken at an airmass of around 1.6 and a PWV-level of about
0.3 mm. Based on ATRAN model calculation of the atmospheric
transmittance vs. PWV for the 18 µm filter we estimated that the
derived Q-band flux for 2005 YU55 must be about 5-10% too
low. We increased the derived 18.7 µm flux of the second night
by 8% and gave a 10% absolute flux calibration error (instead of
7%) to compensate for the additional source of uncertainty. The
final calibrated flux densities are given in Table 2.
2.2. Spacefar-infraredobservationswithHerschel-PACS
The far-infrared observations with the Herschel space obser-
vatory were reported by Muller et al. (2011b). 2005 YU55
crossed the entire visibility window (∼60◦ to ∼115◦ solar elon-
gation) in about 16 hrs and its apparent motion was between
2.8 and 3.8 ◦/h, far outside the technical tracking limit of the
satellite. Therefore, we performed two standard scan-map ob-
servations of 240 s length each -one in the 70/160 µm (2011-
Nov-10 14:52-14:56 UT, OBSID 1342232729) and one in the
100/160 µm filter combination (2011-Nov-10 14:57-15:01 UT,
OBSID 1342232730)- at fixed times at pre-calculated positions
on the sky. Each scan-map consisted of 4 scan-legs of 14 arcmin
length and separated by 4 arcsec parallel to the apparent motion
of the target and with a scan-speed of 20′′/s. During both scan-
map observations 2005 YU55 crossed the observed field-of-view
and the target was seen in each scan-leg. Figure 2 (top) shows
the sky-projected image of the 70 µm band observations. The
PACS photometer takes data frames with 40 Hz, but binned on-
board by a factor of 4 before downlink. We re-centered/stacked
all frames where the satellite was scanning with constant speed
(about 1700 frames in each of the two dual-band measurements)
on the expected position of 2005 YU55. The results are shown
in Fig. 2 (bottom). This technique worked extremely well and
one can clearly see many details of the tripod-dominated point-
spread-function. We performed aperture photometry on the final
calibrated images and estimated the flux error via photometry on
artificially implemented sources in the clean vicinity around our
target. The fluxes were finally corrected for colour terms to ob-
tain monochromatic flux densities at the PACS reference wave-
lengths. These corrections are due to the differences in spectral
energy distribution between 2005 YU55 and the assumed con-
stant energy spectrum ν Fν = const. in the PACS calibration
scheme. The colour-corrections for objects in the temperature
range of ≈ 250 - 400 K are 1.01, 1.03, 1.06 (± 0.01) in blue,
green, red band respectively9. The photometric error of the ar-
tificial sources were combined quadratically with the absolute
flux calibration errors (5% in all 3 bands based on the model un-
certainties of the fiducial stars used in the PACS photometer flux
calibration scheme) and the error related to the colour-correction
(1%). The final monochromatic flux densities and their absolute
flux errors at the PACS reference wavelengths 70.0, 100.0 and
160.0 µm are listed in Table 3.
Fig. 2. Top: Sky-projected PACS image of 2005 YU55 at 70 µm.
Each of the 4 scan-legs has seen the target at a different position.
Bottom: object-centered images of the target in the 3 filters: blue
(70 µm), green (100 µm), red (160 µm). The tripod-dominated
point-spread-function is clearly visible.
2.3. GroundbasedmillimeterobservationswiththeSMA
We performed observations of 2005 YU55 a few hours past clos-
est Earth approach on Nov. 9, 2011 using the Submillimeter
Array (SMA) located near the summit of Mauna Kea in Hawaii.
The SMA was operated in separated sideband mode with
2 GHz continuum bandwidth per sideband. The lower sideband
was tuned to 220.596 GHz and the upper sideband (USB) at
230.596 GHz, providing a mean frequency of 225.596 GHz,
or 1328.9 µm (covering the range from 1300.1 to 1359.0 µm).
Complex gains were obtained from several different quasars as
the asteroid moved across the sky. The amplitude scale was
corrected for Earth atmospheric opacity through standard sys-
tem temperature calibration, and then corrected to the abso-
lute (Jansky) scale by referencing to observations of Uranus
and Callisto, astronomical sources with flux densities known to
within ∼5% at this frequency.
The measurements were difficult due to poor weather, par-
ticularly atmospheric phase stability, and were further hampered
by the exceptionally rapid motion of the object. The asteroid's
apparent position at its fastest changed by ∼7′′/s relative to side-
real, which is significantly faster than the SMA phase tracking
system (the digital delay software, or DDS) was designed for. To
compensate, a special version of the DDS was created which at-
tempted to track the phase on much shorter timescales. However,
this was only partly successful and there were obvious signs of
decorrelation (loss of signal caused by the motion of the source
relative to the tracked phase center) on most baselines. This re-
quired extensive data flagging and secondary self-calibration of
the amplitude, which introduced significant systematic error in
the flux density scale.
8 Precipitable Water Vapour: this is the main source of opacity at mid-
infrared wavelengths.
9 PACS report PICC-ME-TN-038:
http://herschel.esac.esa.int/twiki/pub/Public/PacsCalibrationWeb/cc report v1.pdf
3
Muller et al.: MIR Observations of (308635) 2005 YU55
Table 2. Observing geometries (miniTAO-centric) and final calibrated flux densities. Negative phase angles: after opposition (object
was trailing the Sun). An absolute flux calibration of 5% (N-band) and 7/10% (1st/2nd day Q-band) has been added. The second day
Q-band data point has been corrected for airmass/PWV effects (see text).
Julian Date
mid-time
λre f
[µm]
FD FDerr
[Jy]
[Jy]
rhelio
[AU]
∆obs
[AU]
α Observatory/
[deg]
Instrument
2455874.46285
2455874.46632
2455874.46979
2455874.47326
2455874.47674
2455874.48021
2455874.48368
2455874.48715
2455874.49062
2455874.49410
2455874.49757
2455874.50104
2455874.50451
2455874.50799
2455874.51146
2455874.52187
2455874.52535
2455874.52882
2455874.54965
2455874.55312
2455874.55660
2455874.56007
2455874.56354
2455874.56701
2455874.57049
2455874.57396
2455874.57743
2455875.51458
2455875.57639
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
18.7
8.9
8.9
8.9
12.2
12.2
12.2
18.7
18.7
18.7
18.7
18.7
18.7
18.7
8.9
189.53
192.82
192.81
196.73
194.37
203.01
204.71
212.57
217.57
223.70
219.77
222.79
222.34
227.06
223.24
126.81
123.50
124.36
225.22
221.94
215.43
261.49
253.99
248.50
247.40
241.90
241.30
28.08
19.60
30.44
30.25
29.53
32.07
33.12
32.83
35.25
34.45
33.65
34.04
32.97
32.28
34.22
34.94
35.74
15.22
13.45
14.10
24.78
24.94
23.59
37.68
33.16
34.51
33.88
36.31
33.92
5.90
2.32
0.9904028
0.9904293
0.9904558
0.9904823
0.9905088
0.9905352
0.9905617
0.9905882
0.9906147
0.9906411
0.9906676
0.9906941
0.9907206
0.9907471
0.9907735
0.9908530
0.9908794
0.9909059
0.9910648
0.9910912
0.9911177
0.9911442
0.9911706
0.9911971
0.9912236
0.9912501
0.9912765
0.9983930
0.9988611
0.0021426484
0.0021415786
0.0021408565
0.0021404827
0.0021404572
0.0021407803
0.0021414518
0.0021424716
0.0021438391
0.0021455542
0.0021476157
0.0021500229
0.0021527749
0.0021558704
0.0021593083
0.0021716598
0.0021764507
0.0021815751
0.0022191984
0.0022265916
0.0022342979
0.0022423147
0.0022506389
0.0022592671
0.0022681963
0.0022774233
0.0022869445
0.0084376885
0.0089028990
-97.17 miniTAO/MAX38a
-96.46 miniTAO/MAX38a
-95.74 miniTAO/MAX38a
-95.03 miniTAO/MAX38a
-94.32 miniTAO/MAX38a
-93.61 miniTAO/MAX38b
-92.89 miniTAO/MAX38b
-92.18 miniTAO/MAX38b
-91.47 miniTAO/MAX38b
-90.76 miniTAO/MAX38b
-90.05 miniTAO/MAX38c
-89.34 miniTAO/MAX38c
-88.63 miniTAO/MAX38c
-87.93 miniTAO/MAX38c
-87.22 miniTAO/MAX38c
-85.13 miniTAO/MAX38d
-84.43 miniTAO/MAX38d
-83.74 miniTAO/MAX38d
-79.67 miniTAO/MAX38e
-79.00 miniTAO/MAX38e
-78.34 miniTAO/MAX38e
-77.68 miniTAO/MAX38 f
-77.03 miniTAO/MAX38 f
-76.38 miniTAO/MAX38 f
-75.74 miniTAO/MAX38 f
-75.10 miniTAO/MAX38 f
-74.47 miniTAO/MAX38 f
-19.23 miniTAO/MAX38g
-18.57 miniTAO/MAX38h
Notes. (a,b,c,d,e, f ,g,h) for the χ2 analysis in Section 3 we used the mean fluxes of each group for calculation efficiency reasons.
Table 3. Observing geometries (Herschel-centric) and final calibrated flux densities. Negative phase angles: after opposition.
Julian Date
mid-time
2455876.120565
2455876.120565
2455876.124075
2455876.124075
λre f
[µm]
70.0
160.0
100.0
160.0
FD FDerr
[Jy]
[Jy]
rhelio
[AU]
∆obs
[AU]
α Observatory/
[deg]
Instrument
12.35
2.55
6.87
2.66
0.63
0.13
0.35
0.14
1.002978
1.002978
1.003004
1.003004
0.005403
0.005403
0.005415
0.005415
-70.88 Herschel-PACS
-70.88 Herschel-PACS
-70.62 Herschel-PACS
-70.62 Herschel-PACS
Table 4. Observing geometries (SMA-centric) and final calibrated flux densities. Negative phase angles: after opposition.
Julian Date
mid-time
λre f
[µm]
FD FDerr
[Jy]
[Jy]
rhelio
[AU]
∆obs
[AU]
α Observatory/
[deg]
Instrument
2455874.95042
1328.9
0.075
0.020
0.9941165
0.0042883
-34.66
SMA/230 GHz receiver
Despite these challenges, we obtained a clear detection of
the object. Figure 3 shows the 1.3 mm image of 2005 YU55 af-
ter both phase reference calibration and further self-calibration.
The target itself was unresolved and the oblong image of the as-
teroid in Figure 3 is simply due to the PSF10 of the instrument
for the observations, which is shown in the lower left corner as
10 Point Spread Function
4
an ellipse of 3.73′′ × 2.58′′ in size, with a major axis position
angle of 83.66◦ East of North. Over the 3.5 hours of observation
(UT Nov 9, 2011 09.16 - 12.46 hrs) we further see the expected
drop in flux density as the source recedes, consistent with the
apparent size decrease with time. While the detection is of high
signficance (SNR ∼35), the systematic problems of compensat-
ing for the tracking-induced decorrelation along with the poor
weather dominated the flux-density error budget. Taking all ef-
Muller et al.: MIR Observations of (308635) 2005 YU55
fects into account we obtained a flux density of 75 ± 25 mJy at
observation mid-time (UT Nov 9, 2011 10:49, see Tbl. 4).
Fig. 3. SMA image of 2005 YU55 at 1.3 mm. The ellipse repre-
sents the 2-dimensional full width at half maximum of the syn-
thesized beam of the array, which is effectively the PSF of the
instrument for the observations of an unresolved target.
2.4. Auxiliarydatasets
Radar measurements. Nolan et al. (2010), Busch et al. (2012)
and Taylor et al. (2012a; 2012b) presented results obtained
by radar measurements using the Arecibo S-band, the Deep
Space Network Goldstone DSS-14 and DSS-13, Green Bank
Telescope and Arecibo/VLBA (radar speckle tracking). They
found 2005 YU55 to be a dark (at radar/radio wavelengths),
spherical object of about 400 m diameter and a rotation period of
roughly 18 hrs (Nolan et al. 2010; Taylor et al. 2012a). Busch et
al. (2012) confirmed the nearly spheroidal shape and determined
the maximum dimensions of the object to be 360 ± 40 m in all
directions. The radar team estimated the pole direction from the
motion of the radar speckle pattern during three days of observa-
tions after the flyby. Combining the radar images and the speckle
data excluded all prograde pole directions, and restricted the pos-
sible retrograde poles to (λecl, βecl) = (20◦, -74◦) ± 20◦ with a
rotation period of 19.0 ± 0.5 hrs and consistent with a principle-
axis rotation.
Thermal infrared observations from Gemini-North/Michelle.
Lim et al. (2012a; 2012b) obtained thermal infrared photometry
and spectroscopy in N- and Q-band using the Michelle instru-
ment at Gemini-North. According to their thermal model analy-
sis (Tss = 360 - 370 K; η ≈ 1.25-1.5) the thermal measurements
are consistent with an object diameter of 400 m, but the best fit
to their data was found for a size of 322 ± 18 m and a maxi-
mum subsolar temperature Tss of 409 ± 12 K (thermal model η
≈ 0.93). More recently, Lim et al. (2012c) combined their ther-
mal data with results from radar measurements and find now an
equatorial diameter of 380 ± 20 m and a thermal inertia Γ ≈ 500
- 1500 Jm−2s−0.5K−1. They also calculated values for the effec-
tive diameter via radiometric techniques and based only on their
thermal data. They found an effective diameter of 310 m for a
low thermal inertia of 350 Jm−2s−0.5K−1 and of 350 m for a ther-
mal inertia of 1000 Jm−2s−0.5K−1 (DPS meeting #44, #305.01
presentation).
Keck adaptive-optics (AO) imaging. Merline et al. (2011;
2012) reported on adaptive optics (AO) imaging of 2005 YU55
during its close fly-by on 2011 Nov 9 UT with the Keck II AO
system NIRC2. The preliminary results were derived under the
assumption of a smooth triaxial ellipsoid having a principle-
axis rotation of 18 hr. They found a preference for poles in the
southern sub-latitudes and an effective object diameter of 307
± 15 m. This would be consistent with the radar-favoured ret-
rograde sense of rotation meaning that the object presented a
warm terminator during its close approach. In addition, they give
two explicit solutions: (a) prograde pole with (λecl, βecl) = (339◦,
+84◦) ± 6◦ and object dimensions of 337 × 324 × 267 m (± 15 m
in each dimension), corresponding to a spherical equivalent di-
ameter of 308 ± 9 m; (b) retrograde pole with (λecl, βecl) = (22◦,
-35◦) ± 15◦ and object dimensions of 328 × 312 × 245 m (±
15, 15, 30 m) corresponding to a spherical equivalent diameter
of 293 ± 14 m.
VLT-NACO speckle imaging observations. Sridharan et al.
(2012) performed VLT-NACO speckle imaging in Ks band in
no-AO mode. The observations on 2005 YU55 were carried out
one hour (10-min block) and two hours (15-min block) after the
closest Earth approach, interleaved by sky background and cali-
bration observations. The planned closed-loop AO observations
failed due to poor observing conditions and only no-AO mode
(speckle imaging mode) observations were possible. They found
that 2005 YU55 has a spherical shape with a mean diameter of
about 270 m. At the same time they extracted a mean diameter of
261±20 m × 310±30 m from edge-enhanced image reconstruc-
tions. The large uncertainties are due to the theoretical resolution
of 95 m at the distance of the object and the final image quality.
CCD photometric observations. CCD lightcurve measure-
ments from different observers were analysed by Warner et al.
(2012a; 2012b). Their analysis resulted in two possible synodic
periods of 16.34 ± 0.01 h with an amplitude of 0.24 ± 0.02 mag
(9-17 Nov, 2011) and 19.31 ± 0.02 h with an amplitude of 0.20
± 0.02 mag. The first one was apparently supported by the ini-
tial radar analysis, while the second one is now the currently
favoured solution by the radar team. The 19.31 h lightcurve has
a bimodal shape and there seem to be indications for a non-
principal axis rotation. Due to a large phase angle coverage of
the CCD data they were also able to derive the absolute R-band
magnitude HR = 20.887 ± 0.042 and the phase slope parameter
G = -0.147 ± 0.014. With an assumed V-R value of 0.38 they
calculated the absolute V-band magnitude HV = 21.27 ± 0.05.
Absolute magnitude and phase curve. Based on Bessel R-
band photometry and long-slit CCD spectrograms during the
2010 and 2011 apparitions, Hicks et al. (2010; 2011) reported
an absolute R-band magnitude of HR = 20.73 and a phase slope
parameter G = -0.12 describing a very steep phase curve which
is typically found for low-albedo C- and P-type asteroids. They
measured a V-R colour of 0.37 mag leading to an absolute V-
band magnitude of HV = 21.1 ± 0.1. An indpendent work by
5
Muller et al.: MIR Observations of (308635) 2005 YU55
Bodewits et al. (2011) presented a V-band absolute magnitude
of HV = 21.2 when applying a phase curve derived from UV
measurements (GUV = -0.13).
Table 5. Summary of general TPM input parameters and applied
ranges.
3. Thermophysical model analysis
For the analysis of our thermal data (miniTAO/MAX38, SMA,
Herschel/PACS) we applied a thermophysical model (TPM)
which is based on the work by Lagerros (1996; 1997; 1998).
This model is frequently and successfully applied to near-Earth
asteroids (e.g., Muller et al. 2004; Muller et al. 2005; Muller et
al. 2011a; Muller et al. 2012), to main-belt asteroids (e.g., Muller
& Lagerros 1998; Muller & Blommaert 2004), and also to more
distant objects (e.g. Horner et al. 2012; Lim et al. 2010). The
TPM takes into account the true observing and illumination ge-
ometry for each observational data point, a crucial aspect for the
interpretation of our 2005 YU55 observations which cover a wide
range of phase angles. The TPM allows to specify a shape model
and spin-vector properties. The heat conduction into the surface
is controlled by the thermal inertia Γ. The observed mid- and far-
IR fluxes are connected to the hottest regions on the asteroid sur-
face and dominated by the diurnal heat wave. The seasonal heat
wave is less important and therefore not considered here. The
infrared beaming effects are calculated via a surface roughness
model, described by segments of hemispherical craters. Here,
mutual heating is included and the true crater illumination and
the visibility of shadows is considered. The level of roughness
is driven by the r.m.s. of the surface slopes which correspond
to a given crater depth-to-radius value combined with the frac-
tion of the surface covered by craters, see also Lagerros (1996)
for further details. We used a constant emissivity of 0.9 at all
wavelengths, knowing that the emissivity can decrease beyond
∼200 µm in some cases (e.g., Muller & Lagerros 1998; 2002).
All of our data -except the SMA data point which has a large
errorbar- have been taken at wavelength <200 µm and the con-
stant emissivity is therefore a valid assumption. The TPM input
parameters and applied variations are listed in Table. 5.
3.1. Usingasphericalshapemodel
We started our analysis with a spherical shape model to see
which spin-axis orientations, sizes, geometric albedos, and ther-
mal properties produce acceptable solutions with reduced χ2-
values11 around or below 1.0. For the spin-axis solutions we used
all values specified in literature and many additional orientations
to cover the entire λecl-βecl space. For the calculation of the re-
duced χ2-curves we consider the true observing and illumina-
tion constellation (helio-centric and observer-centric distances,
phase angle, spin-axis orientation) for each epoch and then we
compare with the corresponding measurement. These calcula-
tions are done for a wide range of thermal inertias and different
levels of surface roughness as specified in Table 5. An exam-
ple for the application of this technique can be found in Muller
et al. (2011a). Each model setup produces a curve of reduced
χ2-values as a function of thermal inertia. Figure 4 shows these
curves for all different spin-axis orientation, a rotation period of
19.31 h and an intermediate level of surface roughness (r.m.s. of
surface slopes of 0.3). Reduced χ2-values around or below 1.0
correspond to TPM solutions which explain all observed fluxes
11 reduced χ2-values were calculated via χ2
reduced = 1/(N-ν) P ((obs-
mod)/err)2, with ν being the number of free degrees of freedom; here
ν=2 since we solve for diameter and thermal inertia; obs is the observed
and mod the model flux, err the absolute photometric error.
6
Param.
Value/Range
Remarks
Γ
ρ
f
ǫ
HV-mag.
G-slope
shape
Psid [h]
spin-axis
(λecl,βecl)
0...3000
0.1...0.8
0.6a
0.9b
21.2 ± 0.15 mag
-0.13 ± 0.02
spherical/ellipsoidal
16.34 h; 19.31 h
J m−2 s−0.5 K−1, thermal inertia
(25 values spread in log-space)
r.m.s. of surface slopes, steps of 0.1
surface frac. covered by craters
λ-independent emissivity
average of published values
average of published values
info from radar and AO
Warner et al. (2012a; 2012b)
(20.0◦, -74.0◦) ± 20◦
(339.0◦, +84.0◦) ± 6◦ Merline et al. (2011; 2012)
(22.0◦, -35.0◦) ± 15◦ Merline et al. (2011; 2012)
Busch et al. (2012)
(309.3◦, +89.5◦)c
(129.3◦, -89.5◦)d
(337.2◦, -13.9◦)
(157.2◦, +13.9◦)
(273.0◦, +1.7◦)
(93.0◦, -1.7◦)
(337.2◦, +76.1◦)
(157.2◦, -76.1◦)
(273.0◦, -88.3◦)
(93.0◦, +88.3◦)
(0/90/180/270◦, ±60◦)
(0/90/180/270◦, ±30◦)
(0/90/180/270◦ , 0◦)
obliquity 0◦ (prograde)
obliquity 180◦ (retrograde)
pole-on case1 for Herschel obs.
pole-on case2 for Herschel obs.
pole-on case1 for TAO/MAX38
pole-on case2 for TAO/MAX38
equ.-on case1 for Herschel obs.
equ.-on case2 for Herschel obs.
equ.-on case1 for TAO/MAX38
equ.-on case2 for TAO/MAX38
intermediate orientations
intermediate orientations
pole in ecliptic plane
Notes. (a) see Lagerros 1998 section 3.3; (b) see text for further details;
(c) spin-axis orientation close to ecliptic north pole; (d) spin-axis orien-
tation close to ecliptic south pole
in a statistically acceptable way. There are several spin-axis ori-
entations which produce an excellent match to all our thermal
measurements at thermal inertia values in the range between ap-
proximately 200 and 1500 Jm−2s−0.5K−1.
The distribution of the reduced χ2-minima along the ecliptic
longitudes and latitudes is shown in Figure 5. There are large
zones in the λecl.-βecl-space which can be excluded with high
probability (light blue, green, yellow, red zones), but there re-
main several possible spin-axis orientations compatible with our
dataset (dark blue zones), including the radar and AO solutions.
Both figures (Figs. 4 & 5) have a slight dependency on
the selected surface roughness (for both figures we have used
r.m.s. of surface slopes of 0.3). In general, lower roughness
(r.m.s. of surface slopes at 0.1) produces lower χ2-minima
and at smaller thermal inertia values going down to about
200 Jm−2s−0.5K−1. Higher values for the surface roughness
(r.m.s. of surface slopes of ≥0.5) shift the χ2-minima to values
well above 1.0 and towards higher thermal inertia going up to
about 1500 Jm−2s−0.5K−1. It is interesting to note that the pro-
grade AO solution (solid line in Fig. 4) works very well (χ2-
minima very close to 1.0) for a low surface roughness, while the
radar solution produces a better match in case of a high surface
roughness.
3.2. Influenceofthespin-axisorientation
As a next step, we investigate the influence of different spin-axis
orientations on the size and albedo solutions. We determined the
Muller et al.: MIR Observations of (308635) 2005 YU55
derived thermal inertias change significantly with roughness at
similar χ2-values, indicating that we cannot resolve the degen-
eracy between roughness and thermal inertia with our dataset.
A smoother surface is connected to lower values for the thermal
inertia, the rougher surfaces require higher thermal inertias.
Fig. 4. Calculation of reduced χ2-values for all specified spin-
axis orientations, a fixed rotation period of 19.31 h and an in-
termediate surface roughness level (r.m.s. of surface slopes of
0.3). The prograde AO solution (solid line) and the radar solu-
tion (dashed line) for the spin-vector are indicated in the figure.
]
g
e
d
[
e
d
u
t
i
t
a
l
c
i
t
p
i
l
c
E
80
60
40
20
0
−20
−40
−60
−80
0
50
100
8
7
6
5
4
3
2
1
250
300
350
150
200
Ecliptic longitude [deg]
Fig. 5. The χ2-minima calculated for all spin-axis orientations
listed in Table 5 and for an intermediate level of roughness
(r.m.s.slope 0.3). The dark blue zones indicate spin-poles which
allow us to obtain an acceptable match to all thermal data simul-
taneously (reduced χ2-values around or below 1.0). The radar
and both AO solutions are indicated by the crossed circles. Note
that the size, albedo and thermal inertia are free parameters and
only the best possible solution for each spin-axis has been con-
sidered.
χ2-minima for all listed spin-axis orientations and for four dif-
ferent levels of roughness (r.m.s. of surface slopes at 0.1, 0.3,
0.5 and 0.8). Figure 6 shows how the corresponding radiometric
sizes and geometric albedos are distributed in the reduced χ2-
picture. We connected the four χ2-minima belonging to the AO-
solution (solid line) and the ones belonging to the radar solution
(dashed line) in Fig. 6. These lines show that the connected size
and albedo values remain stable, just the fit gets better (lower χ2-
minima) for specific roughness settings. We also found that the
Fig. 6. The distribution of the χ2-minima and the related effective
diameter (top) and geometric albedos (bottom). The four differ-
ent levels of roughness are indicated by different symbols. The
values for the prograde AO solution (solid line) and the radar
solution (dashed line) are connected in the figures.
The thermal data are compatible with different spin-axis ori-
entations, but the size, the geometric albedo and also the pos-
sible thermal inertias are very well constrained by our thermal
dataset. The best solutions are found for an effective diameter
of about 310 m, if we include the best solutions for the prograde
AO spin-axis and the radar spin-axis orientations, then the possi-
ble diameter range goes from 295 to 335 m (see Fig. 6, top). For
the geometric albedo we find a value of about 0.062 and a possi-
ble range between 0.053 to 0.067 (see Fig. 6, bottom). Figure 7
shows how our best TPM solution translates the insolation dur-
ing the epoch of the Herschel measurement into a thermal pic-
ture of the surface as seen from Herschel. For the calculations
we used a spin-axis orientation of (λecl, βecl) = (60◦, -60◦) and
a spherical shape model with a total of 800 facets. The large in-
fluence of the thermal inertia in combination with the object's
7
Muller et al.: MIR Observations of (308635) 2005 YU55
rotation is the reason for the warm temperatures also in regions
without direct illumination.
4. Discussions
4.1. Comparisonwiththeradarresults
Insolation [W/m2]
Temperature [K]
1200
1000
800
600
400
200
350
300
250
200
150
100
50
Fig. 7. TPM picture of 2005 YU55 as seen from Herschel on
2011-Nov-10 14:55 UT in the object-centered reference frame
(z-axis along the object's rotation axis) and with the Sun at a
phase angle of -71◦, spin-axis orientation: (λecl, βecl) = (60◦, -
60◦), spherical shape model with a total of 800 facets. Top: inso-
lation in W/m2. Bottom: temperature in K.
8
The comparison between the radar results (Busch et al. 2012)
and our findings is very interesting. If we use the radar diame-
ter (360 ± 40 m, close to a spheroidal shape) and the spin-axis
properties ([λecl, βecl] = [20◦, -74◦] ± 20◦, Psid = 19.0 ± 0.5 h) it
is not possible to find an acceptable match to our thermal mea-
surements. The reduced χ2-minima stay always well above 2.0
and the match between TPM-predictions and observed fluxes is
very poor. Even for the lowest diameter limit of 320 m the model
calculations would exceed the measured fluxes systematically by
15-25%. At a diameter of 360 m the model fluxes are already 30-
40% above the measurements. The radar size estimates are -as
the radiometric size estimates- model dependent. The spin-axis
orientation as well as the rotation rate have a larger influence on
the radar solution (e.g., Ostro et al. 2002) than they have on the
radiometric solution. The radar images are dominated by the sur-
face part which is closest to the antenna while the thermal data
are tidely connected to the entire cross-section at the moment
of observation. This might explain the differences between both
techniques.
However, we do find an acceptable match to all thermal
data if we just use the radar spin-properties combined with a
high level of surface roughness (r.m.s. of surface slopes of 0.8).
But the corresponding diameter is only 299 m -well outside the
radar derived range- with a pV =0.067 and a thermal inertia of
400 Jm−2s−0.5K−1. In fact, all high obliquity cases with βecl ≤ -
60◦ (retrograde sense of rotation) produce small diameters in the
range 300-310 m, while only the low obliquity cases with βecl ≥
+60◦ (prograde sense of rotation) produce larger effective diam-
eters of 325-340 m.
4.2. ComparisonwithAOandspeckleresults
The Keck AO results presented by Merline et al. (2012) compare
better with our findings. Table 6 summarises the AO and our
radiometric results.
Table 6. Comparison between AO results and our findings.
sense of
rotation
prograde
retrograde
retrograde
spin-axis
(λecl, βecl)
AO-size
[m]
339◦, +84◦
22◦, -35◦
southern poles
337×324×267
328×312×245
AO Dequ
[m]
308 ± 9
293 ± 14
307 ± 15
TPM-Dequ
[m]
333
299a
300-310b
Notes. (a) this solution requires an unacceptably high thermal inertia of
well above 2000 Jm−2s−0.5K−1;
(b) diameter range of all high obliquity
cases βecl ≤ -60◦
The southern rotational poles are not specified in detail by
Merline et al. (2012), but here we see for the first time an agree-
ment between the derived sizes. The originally specified retro-
grade pole towards an ecliptic latitude of -35◦ is very unlikely:
acceptable TPM solutions (with reduced χ2-minima below 2.0)
are only found if the thermal inertia would be well above
2000 Jm−2s−0.5K−1, an unrealistically high value which has never
been measured before. It should be noted here that the highest
Muller et al.: MIR Observations of (308635) 2005 YU55
derived thermal inertias are still below 1000 Jm−2s−0.5K−1 (e.g.,
Delbo et al. 2007) and that our mid- to far-IR data originate in
the top layer on the surface. We don't see any signatures of sub-
surface layers where the thermal inertia could be significantly
higher (Keihm et al. 2012).
The speckle observation in no-AO mode presented by
Sridharan et al. (2012) revealed a roughly spheroidal shape with
a mean diameter of 270 m. By using a more sophisticated re-
construction technique they estimated a mean diameter of 261
± 20 m × 310 ± 30 m, corresponding to an object-averaged size
of approximately 285 ± 25 m. Within the errorbars, this value
agrees with our radiometrically derived diameter of 300-310 m
and it also creates doubts if the large radar size is realistic. The
indications for a diameter close to 300 m makes also the various
prograde solutions more unlikely, which all require diameters in
the range 325-340 m.
4.3. Spin-axisproperties
Combining the spin-axis information given by Busch et al.
(2012), Merline et al. (2012) and our findings (see Fig. 4), our
analysis supports a retrograde sense of rotation with a possible
spin-axis orientation of (λecl, βecl) = (60◦ ± 30◦, -60◦ ± 15◦). The
relatively large errors in (λecl, βecl) are covering also the possible
solutions connected to the different roughness levels mentioned
before. If we use this solution, then the size estimate from AO
observations matches our radiometrically derived optimal size
and we also have an agreement with the radar derived spin-pole.
The discrepancy with the radar size remains.
Our thermal observations cover a wide range of phase an-
gles, wavelengths and different illumination and observing ge-
ometries. This allowed us to exclude many spin-axis orientations
(see Fig. 5). Nevertheless, we could not find a strong prefer-
ence for a single spin-axis orientation nor for the sense of ro-
tation. Even very extreme solutions like the retrograde radar so-
lution and the prograde AO solution seem to explain the data
equally well. This is very surprising. Based on our previous
modeling experiences for 1999 JU3 (Muller et al. 2011a) and
1999 RQ36 (Muller et al. 2012) based on much smaller sets of
thermal data, we expected to find a unique spin-axis solution.
But this might be an indication that 2005 YU55 is a tumbler with
a strongly time-dependent orientation of the spin-axis (for fur-
ther details on tumbling asteroids see Pravec et al. 2005). Busch
et al. (2012) speculated already about the possibility that ter-
restrial tides might have torqued the object into a non-principal
axis spin state. However, their observations are consistent with
a principle-axis rotation. Warner et al. (2012b) found two, non-
commensurate solutions for the rotation period (16.34 ± 0.01 h;
19.31 ± 0.02 h) which they could not fully explain. They suggest
that a non-principal axis rotation should be considered. After the
radar and lightcurve analysis, the thermal analysis is now also
pointing towards the possibility of a non-principle axis rotation.
We also looked into the influence of the two published rota-
tion periods. But the ≈3 h difference between the two available
periods did not affect our radiometric solutions significantly.
The longer rotation period is typically requiring slightly higher
inertias to produce the same disk-integrated flux, but this is a
marginal effect here in this case.
4.4. Comparisonwithotherthermalmeasurements
Instead of comparing our TPM radiometric results with the pre-
liminary results produced by Lim et al. (2012a; 2012b; 2012c)
via a simple thermal model, we predicted flux densities for
the epochs and the wavelength bands of the Michelle/Gemini
North observations shown in Figure 2 in Lim et al. (2012b). For
the TPM prediction we simply used our best effective diame-
ter (310 m) and albedo (pV = 0.062) solution connected to our
preferred spin-axis orientation of (λecl, βecl) = (60◦, -60◦). The
thermal inertia and roughness levels are less well constrained
and our dataset does not allow to break the degeneracy between
these two parameters. A low roughness (r.m.s. of surface slopes
= 0.1) combined with small values of the thermal inertia of about
200 Jm−2s−0.5K−1 would explain our measurements as well as
higher roughness levels (r.m.s. of surface slopes = 0.5) com-
bined with higher thermal inertia around 800 Jm−2s−0.5K−1. We
selected an intermediate solution (r.m.s. of surface slopes = 0.3;
thermal inertia = 500 Jm−2s−0.5K−1).
The Gemini-North/Michelle photometry shown in the
Figure 2 in Lim et al. (2012b) was taken on 09-Nov-2011 11:02-
11:15 UT (α = -34.0◦, r = 0.994 AU, ∆ = 0.004 AU) and on
10-Nov-2011 09:32 - 11:52 UT (α = -15.5◦, r = 1.001 AU, ∆
= 0.012 AU). Since the calibrated flux densities and errors are
not explicitly given, we only could do a qualitative compari-
son. Table 7 shows our TPM prediction for both epochs and the
Michelle reference wavelengths in Jansky and W/m2/µm.
Table 7. TPM flux predictions for the Michelle bands and both
observing epochs.
Wavelength
λc [µm]
09-Nov-2011 11:08UT
[W/m2/µm]
[Jy]
10-Nov-2011 10:50UT
[Jy]
[W/m2/µm]
7.9
8.8
9.7
10.3
11.6
12.5
18.5
69.2
83.7
95.3
101.5
110.8
114.6
109.2
3.3e-12
3.2e-12
3.0e-12
2.9e-12
2.5e-12
2.2e-12
1.0e-12
10.6
12.9
14.7
15.6
17.0
17.6
16.6
5.1e-13
5.0e-13
4.7e-13
4.4e-13
3.8e-13
3.4e-13
1.5e-13
Our TPM-predictions agree very well with the observed
fluxes and errorbars presented in Lim et al. (2012b). For the first
epoch we estimated that the agreement is within about 10% at
all wavelengths, while for the second epoch the TPM prediction
seems to be about 5-15% below the observed fluxes.
We also tested the low-roughness/low-inertia case mentioned
before and indeed it produces very similar fluxes and the agree-
ment is on a similar level. The high-roughness/high-inertia case
is less convincing, the TPM predictions are systematically low
by 5-20%. The Michelle/Gemini North data favour a thermal in-
ertia value in the range 200-700 Jm−2s−0.5K−1, combined with
an intermediate to low roughness level (r.m.s. of surface slopes
0.1-0.5), also in agreement with the lowest reduced χ2-values in
Fig. 4.
4.5. Overallfittothemeasurements
We tested the quality of the final solution for 2005 YU55 against
the observed and calibrated flux densities by calculating the
TPM predictions for each of data point listed in Tables 2, 3,
and 4. The observed and calibrated mono-chromatic flux den-
sities are shown in Fig. 8 together with the TPM predictions
9
Muller et al.: MIR Observations of (308635) 2005 YU55
for the specific observing geometries. The observation/TPM ra-
tios are very sensitive to wavelength-dependent effects (related
surface roughness and thermal inertia), phase-angle dependent
effects (a wrong thermal inertia would cause before/after oppo-
sition asymmetries), and shape effects (ratios as a function of
rotational phase). An overall ratio close to 1.0 indicates that the
size and thermal properties (and in second order also albedo) are
correctly estimated. Figure 9 shows how well our final TPM so-
lution explains our thermal data covering a wide range of wave-
lengths from 8.9 µm to 1.3 mm and taken at very different phase
angles ranging from -97◦ to -18◦. No trends with wavelength nor
with phase angle can be seen.
Fig. 8. Observered and calibrated flux densities together with the
corresponding TPM prediction. The model predictions for the
MAX38 data are shown at the start and end time of each observ-
ing day. The distance between observer and target and also the
phase angle were rapidly changing during the close encounter
period of three days. For the PACS data the model prediction
from 5 to 1500 µm is shown.
Figure 9 also shows that 2005 YU55 must be close to a
sphere. An elongated or strangely shaped body would produce
a thermal lightcurve, but our dataset does not show any signifi-
cant deviations at specific rotational phases (bottom figure). But
not all rotational phases have been covered by our thermal mea-
surements and some of the observational errors are large. There
is also the possibility that effects of an ellipsoidal shape could
have been compensated by roughness effects (a larger cross-
section combined with a low surface roughness could produce
the same flux levels as a smaller cross-section combined with
high surface roughness). Figure 9 (bottom) would then also show
a constant ratio at all rotational phases. But since the roughness
influences the flux in a wavelength-dependent manner (see e.g.,
Muller 2002, Fig. 3), one should then see a larger scatter in Fig.
9 (top) at short wavelengths where the roughness has the great-
est influence on the observed fluxes. At long wavelengths (be-
yond ∼20 µm) the effects of roughness are much smaller and
the shape effects are dominating. Shape effects or combined
shape/roughness variations are not seen in our dataset.
We did also an additional test to see if the optical lightcurve
amplitude of 0.20 ± 0.02 mag (Warner et al. 2012a; 2012b) is
compatible with our findings. Such an amplitude would mean
that the flux at lightcurve maximum is about 1.2 times the flux at
lightcurve minimum, which would require a SNR>10 time series
10
Fig. 9. Observered and calibrated flux densities divided by the
corresponding TPM prediction. Top: as a function of wave-
length. Middle: as a function of phase angle. Bottom: as a func-
tion of rotational phase.
Muller et al.: MIR Observations of (308635) 2005 YU55
data set for confirmation. The PACS data are of sufficient qual-
ity, but they are taken at a single epoch. The miniTAO/MAX38
measurements have too large error bars, related mainly to sys-
tematic errors in the absolute flux calibration scheme. However,
we looked at the relative variation of the 22 miniTAO/MAX38
data points taken at 18.7 µm with respect to the spherical shape
model flux predictions. The deviations never exceed 10%, but
these data cover only a very limited range of rotational phases
(from 195 to 245◦ and a single point at 305◦ in the bottom of
figure 9). The thermal data are therefore perfectly compatible
with the optical lightcurve results and there are not indications
for large deviations from a spherical shape.
4.6. Errorcalculations
We combine the constraints from the radar measurements (retro-
grade sense of rotation, estimate of spin-axis orientation), the
AO findings (effective diameter of 307 ± 15 m for "southern
poles"), and the speckle technique (object-averaged diameter of
285 ± 25 m) with the χ2 analysis for the possible spin-axis ori-
entations (see Figs. 4, 5 and corresponding figures for different
roughness levels which are not shown here). For a good fit the
reduced χ2-values should be close to 1 and we estimated for our
dataset that the 3-σ confidence level for the reduced χ2 is around
1.6. This lead to an estimated spin-axis orientation of (λecl, βecl)
= (60◦ ± 30◦, -60◦ ± 15◦).
We can use the 3-σ threshold in reduced χ2 also for the
derivation of the corresponding size and albedo range. Figure 10
shows the size and albedo solutions for the full range of ther-
mal inertias (from 0 to 3000 J m−2 s−0.5 K−1), the four different
levels of roughness (r.m.s.-slopes of 0.1, 0.3, 0.5, 0.8, shown
with different symbols) and for all spin-axis solutions compat-
ible with (λecl, βecl) = (60◦ ± 30◦, -60◦ ± 15◦). Based on the 3-σ
confidence level we derived a possible diameter range of 295 to
322 m, 0.057 to 0.068 for the geometric albedo, and a thermal
inertia larger than 150 J m−2 s−0.5 K−1.
As a second step we looked in more details at the derived
size, albedo and thermal inertia ranges. The solutions close to
the 3-σ threshold in our χ2-analysis are very problematic in the
sense that they produce strong trends in the observation/model
figures (see Fig. 9) either with wavelengths and/or with phase
angle. These kind of trends are very difficult to catch in an auto-
matic χ2-analysis. We therefore moved back to the 1-σ solutions,
corresponding to a possible diameter range of 300-312 m, a geo-
metric albedo range of 0.062-0.067, and a thermal inertia range
of 350-1000 J m−2 s−0.5 K−1. The smallest thermal inertia values
are connected to low roughness values (r.m.s.-slopes ≤ 0.3) and
the largest thermal inertia values to very rough surface levels
(r.m.s.-slopes ≥ 0.5). The calculations for the Michelle/Gemini
North data put another constraint on the thermal inertia and re-
duce the possible range to 350 - 800 J m−2 s−0.5 K−1. The derived
radiometric albedo range of 0.062-0.067 is connected to the HV
magnitude of 21.2 m, if we include the ±0.15 mag, then the pos-
sible range is significantly bigger: from 0.055 to 0.075.
5. Conclusions
Here is a short summary of our findings for the near-Earth aster-
oid 2005 YU55:
1. Our thermal data can be explained via a spherical shape
model without seeing significant offsets at specific rotational
phases, showing that 2005 YU55 is almost spherical.
Fig. 10. The size and albedo solutions for the full range of ther-
mal inertias, the four different levels of roughness and for the
most likely spin-axis solutions.
2. Our best spin-axis solution can be specified by (λecl, βecl) =
(60◦ ± 30◦, -60◦ ± 15◦). However, the analysis of the ther-
mal data alone would also allow for specific spin-axis orien-
tations in the northern ecliptic hemisphere with a prograde
rotation of the object.
3. The radiometric analysis of our thermal data which span a
wide range of phase angles and wavelengths (best visible in
the χ2-picture in Fig. 5) is compatible with changing spin-
axis orientations, which might be an indication for a non-
principal axis rotation of 2005 YU55.
4. 2005 YU55 has a possible effective diameter range of Dequ
= 300 - 312 m (equivalent diameter of an equal volume
sphere); this range was derived under the assumption that
the spin-axis is indeed as specified above.
5. The analysis of all available data combined revealed a dis-
crepancy with the radar-derived size.
6. The geometric visual albedo pV was radiometrically derived
to be in the range 0.062 to 0.067 (HV = 21.2 mag) or 0.055 -
0.075 if we include the ±0.15 mag error in HV, in agreement
with the C-type taxonomic classification.
7. 2005 YU55 has a thermal
inertia in the range 350-
800 Jm−2s−0.5K−1, very similar to the value found for the
rubble-pile asteroid (25143) Itokawa by Muller et al. (2005).
We expect therefore that the surface of 2005 YU55 looks
11
Muller et al.: MIR Observations of (308635) 2005 YU55
Vodniza, A.Q. & M.R. Pereira, 2010, DPS meeting #42, #13.25, BAAS 42, 1057
Warner, B. D., Stephens, R. D., Brinsfield, J. W. et al. 2012, The Minor Planet
Bulletin, Association of Lunar and Planetary Observers, Vol. 39, No. 2, 84-
85
Warner, B. D., Stephens, R. D., Brinsfield, J. W. et al. 2012, Asteroids, Comets,
Meteors 2012, conference proceedings, LPI Contribution No. 1667, id. 6013
Yoshii, Y., Aoki, T., Doi, M. et al. 2010, Proc. SPIE 7733, Ground-based and
Airborne Instrumentation for Astronomy III, 773308
also very similar and is composed of low conductivity fine
regolith mixed with larger rocks and boulders which have
much higher thermal inertias.
8. The observed thermal emission can be best reproduced when
considering a low to intermediate roughness with an r.m.s.-
slope of 0.1-0.3; the lower roughness (or smoother surface)
is connected to the lower thermal inertias, while a higher
roughness would require also the higher inertia values.
Acknowledgements. We would like to thank the Herschel operations team which
supported the planning and scheduling of our fixed-time observations. Without
their dedication and enthusiasm these measurements would not have been pos-
sible. The Submillimeter Array is a joint project between the Smithsonian
Astrophysical Observatory and the Academia Sinica Institute of Astronomy and
Astrophysics and is funded by the Smithsonian Institution and the Academia
Sinica. SH is supported by the Space Plasma Laboratory, ISAS/JAXA. AP is
supported by the Hungarian grant LP2012-31/2012.
References
Asano, K., Miyata, T., Sako, S. et al., 2012, Proc. of SPIE, 8446, 115
Bodewits, D., Campana, S., Kennea, J. et al. 2011, Central Bureau Electronic
Telegrams, 2937, 1 (2011)
Busch, M. W., Benner, L. A. M., Brozovic, M. et al. 2012, Asteroids, Comets,
Meteors 2012, conference proceedings, LPI Contribution No. 1667, id. 6179
Cohen, M., Walker, R. G., Carter, B. et al., 1999, AJ, 117, 1864
Delbo, M., dell'Oro, A., Harris, A. W. et al. 2007, Icarus 190, 236
Hicks, M., Lawrence, K., Benner, L. 2010, The Astronomer's Telegram, 2571
Hicks, M., Somers, J., Truong, T. & Teague, S. 2011, The Astronomer's
Telegram, 3763
Horner, J., Muller, T. G., Lykawka, P. S. 2012, MNRAS 423, 2587-2596
Keihm, S., Tosi, F., Kamp, L. et al. 2012, Icarus 221, 395
Lagerros, J. S. V. 1996, A&A 310, 1011
Lagerros, J. S. V. 1997, A&A 325, 1226
Lagerros, J. S. V. 1998, A&A 332, 1123
Lim, T. L., Stansberry, J., Muller, T. G. et al. 2010, A&A 518, 148-152
Lim, L. F., Emery, J. P., Moskovitz, N. A., Granvik, M. 2012, 43rd Lunar and
Planetary Science Conference, LPI Contribution No. 1659, id. 2202
Lim, L. F., Emery, J. P., Moskovitz, N. A., Granvik, M. 2012, Asteroids, Comets,
Meteors 2012, conference proceedings, LPI Contribution No. 1667, id. 6295
Lim, L. F., Emery, J. P., Moskovitz, N. A., et al. 2012, DPS meeting #44, #305.01
Merline, W. J., Drummond J. D., Tamblyn, P. M. et al. 2011, IAU Circular 9242,
http://www.cbat.eps.harvard.edu/iauc/09200/09242.html
Merline, W. J., Drummond J. D., Tamblyn, P. M. et al. 2012, Asteroids, Comets,
Meteors 2012, conference proceedings, LPI Contribution No. 1667, id. 6372
Miyata, T., Sako, S., Nakamura, T. et al., 2008, Proc. SPIE 7014, Ground-based
and Airborne Instrumentation for Astronomy II, 701428
Muller, T. G. & Lagerros, J. S. V. 1998, A&A, 338, 340-352
Muller, T. G. 2002, M&PS, 37, 1919
Muller, T. G. & Blommaert, J. A. D. L. 2004, A&A, 418, 347-356
Muller, T. G., Sterzik, M. F., Schutz, O. et al. 2004, A&A, 424, 1075-1080
Muller, T. G., Sekiguchi, T., Kaasalainen, M. et al. 2005, A&A, 443, 347-355
Muller, T. G., Durech, J., Hasegawa, S. et al. 2011, A&A, 525, 145
Muller, T. G., Altieri, B. & Kidger, M. 2011,
IAU Circular 9241,
http://www.cbat.eps.harvard.edu/iauc/09200/09241.html
Muller, T. G., O'Rourke, L., Barucci, A. M. et al. 2012, A&A, 548, 36-45
Nakamura, T., Miyata, T., Sako, S. et al., 2010, Proc. SPIE 7735, Ground-based
and Airborne Instrumentation for Astronomy III, 773561
Nolan, M. C., Vervack, R. J., Howell, E. S. et al. 2010, American Astronomical
Society, DPS meeting #42, #13.19, Bulletin of the American Astronomical
Society, Vol. 42, p. 1056
Ostro, S., Hudson, R. S., Benner, L. A. M. et al. 2002, in Asteroids III, W. Bottke,
A. Cellino, P. Paolicchi and R. P. Binzel (eds), 151-168
Pravec, P., Harris, A. W., Scheirich, P. et al. 2005, Icarus 173, 108-131
Sako, S., Aoki, T., Doi, M. et al., 2008, Proc. SPIE 7012, Ground-based and
Airborne Telescopes II, 70122T
Somers, J.M. Hicks, M., Lawrence, K. et al., 2010, DPS meeting #42, #13.16,
BAAS 42, 1055
Sridharan, R., Girard, J. H. V., Lombardi, G., Ivanov, V. D., Dumas, C. 2012,
Optical and Infrared Interferometry III. Proceedings of the SPIE, Volume
8445
Taylor, P. A., Nolan, M. C., Howell, E. S. et al. 2012, American Astronomical
Society, AAS Meeting #219, #432.11
Taylor, P. A., Howell, E. S., Nolan, M. C. et al. 2012, Asteroids, Comets, Meteors
2012, conference proceedings, LPI Contribution No. 1667, id. 6340
12
|
1911.00278 | 3 | 1911 | 2019-11-11T08:36:22 | Importance of Giant Impact Ejecta for Orbits of Planets Formed during the Giant Impact Era | [
"astro-ph.EP",
"astro-ph.SR"
] | Terrestrial planets are believed to be formed via giant impacts of Mars-sized protoplanets. Planets formed via giant impacts have highly eccentric orbits. A swarm of planetesimals around the planets may lead to eccentricity damping for the planets via the equipartition of random energies (dynamical friction). However, dynamical friction increases eccentricities of planetesimals, resulting in high velocity collisions between planetesimals. The collisional cascade grinds planetesimals to dust until dust grains are blown out due to radiation pressure. Therefore, the total mass of planetesimals decreases due to collisional fragmentation, which weakens dynamical friction. We investigate the orbital evolution of protoplanets in a planetesimal disk, taking into account collisional fragmentation of planetesimals. For 100 km-sized or smaller planetesimals, dynamical friction is insignificant for eccentricity damping of planets because of collisional fragmentation. On the other hand, giant impacts eject collisional fragments. Although the total mass of giant impact ejecta is 0.1-0.3 Earth masses, the largest impact ejecta are ~ 1,000 km in size. We also investigate the orbital evolution of single planets with initial eccentricities 0.1 in a swarm of such giant impact ejecta. Although the total mass of giant impact ejecta decreases by a factor of 3 in 30 Myrs, eccentricities of planets are damped down to the Earth level (~0.01) due to interaction with giant impact ejecta. Therefore, giant impact ejecta play an important role for determination of terrestrial planet orbits. | astro-ph.EP | astro-ph | Draft version June 15, 2021
Preprint typeset using LATEX style emulateapj v. 01/23/15
9
1
0
2
v
o
N
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
3
v
8
7
2
0
0
.
1
1
9
1
:
v
i
X
r
a
IMPORTANCE OF GIANT IMPACT EJECTA FOR ORBITS OF PLANETS FORMED DURING THE GIANT
IMPACT ERA
Hiroshi Kobayashi, Kazuhide Isoya, and Yutaro Sato
Department of Physics, Nagoya University, Nagoya, Aichi 464-8602, Japan
Draft version June 15, 2021
ABSTRACT
Terrestrial planets are believed to be formed via giant impacts of Mars-sized protoplanets. Planets
formed via giant impacts have highly eccentric orbits. A swarm of planetesimals around the planets
may lead to eccentricity damping for the planets via the equipartition of random energies (dynami-
cal friction). However, dynamical friction increases eccentricities of planetesimals, resulting in high
velocity collisions between planetesimals. The collisional cascade grinds planetesimals to dust un-
til dust grains are blown out due to radiation pressure. Therefore, the total mass of planetesimals
decreases due to collisional fragmentation, which weakens dynamical friction. We investigate the or-
bital evolution of protoplanets in a planetesimal disk, taking into account collisional fragmentation
of planetesimals. For 100 km-sized or smaller planetesimals, dynamical friction is insignificant for ec-
centricity damping of planets because of collisional fragmentation. On the other hand, giant impacts
eject collisional fragments. Although the total mass of giant impact ejecta is 0.1-0.3 Earth masses,
the largest impact ejecta are ∼ 1, 000 km in size. We also investigate the orbital evolution of single
planets with initial eccentricities 0.1 in a swarm of such giant impact ejecta. Although the total mass
of giant impact ejecta decreases by a factor of 3 in 30 Myrs, eccentricities of planets are damped down
to the Earth level (∼ 0.01) due to interaction with giant impact ejecta. Therefore, giant impact ejecta
play an important role for determination of terrestrial planet orbits.
Keywords: Planet formation (1241), Solar system formation (1530), Inner planets (797)
1.
INTRODUCTION
In the standard scenario for terrestrial planet forma-
tion, Mars-sized protoplanets are formed prior to the gas
depletion of the protoplanetary disk (in several Myrs)
with large orbital separations ∼ 10 mutual Hill radii
(e.g., Kobayashi & Dauphas 2013). The gas depletion
triggers the long term orbital instability of protoplan-
ets and the collisions between protoplanets induced by
the orbital instability result in the formation of Earth
or Venus sized planets, which is called the giant impact
stage (Chambers, & Wetherill 1998; Iwasaki et al. 2001).
Most of the masses of the terrestrial planets in the
Solar System is in Earth and Venus, which have low ec-
centricities of 0.017 and 0.007, respectively. The largest
planets formed in orbital simulations for giant impact
stages have much greater eccentricities and inclinations
than those of Earth or Venus (Chambers 2001; Kokubo et
al. 2006). Those eccentricities and inclinations are possi-
ble to be damped via the equipartition of random ener-
gies (dynamical friction) with surrounding planetesimals
(O'Brien et al. 2006; Raymond et al. 2009; Morishima et
al. 2010). However, the surface density of surrounding
planetesimals decreases via the collisional cascade of the
planetesimals (Kobayashi & Tanaka 2010), which reduces
the efficiency of dynamical friction. Therefore, collisional
fragmentation plays an important role in this issue.
On the other hand, a series of giant impacts eject frag-
ments with a total mass comparable to Earth,
result-
ing in the increase of infrared emission. Therefore, gi-
ant impact ejecta explain infrared excesses caused by
warm debris disks around 1AU (Genda et al. 2015), while
cold debris disks beyond 10 AU may be related to col-
[email protected]
lisional fragmentation in planetesimal disks induced by
planet formation (Kobayashi, & Lohne 2014).
Such
giant-impact-ejecta disks may affect the orbital evolu-
tion of protoplanets. The evolution of total masses of gi-
ant impact ejecta is controlled by the collisional cascade.
Therefore, we need to consider the orbital evolution and
the collisional cascade simultaneously.
The orbital evolution of protoplanets in the giant im-
pact stage is mainly treated by N -body simulation. How-
ever, all fragments produced via collisional fragmenta-
tion cannot be treated individually by N body simula-
tion because of computational limitation. Therefore, one
applies the super-particle approximation where a super
particle represents a large number of planetesimals and
fragments. This method is applied for planet formation
(Levison et al. 2012; Morishima 2015; Walsh, & Levison
2019) and for debris disks (Kral et al. 2013; Nesvold et
al. 2013). In this paper, we newly develop the N body
code including the mass evolution of planetesimals via
collisional cascade, which allows us to evaluate dynam-
ical friction and collisional fragmentation in the giant
impact stage. In §. 2, we explain the method to develop
the code. In §. 3, we conduct test calculations for the
collisional cascade and validate the method. In §. 4, we
perform simulations for the orbital evolution of proto-
planets in planetesimal or giant-impact-ejecta disks us-
In §. 5, we discuss the
ing the newly developed code.
effect of remnant planetesimals and giant impact ejecta
on the orbital evolution of protoplanets in the giant im-
pact stage. We summarize our finding in §. 6.
2. METHOD
In the giant impact stage, the orbital evolution and
collisions of protoplanets occur. The gravitational inter-
2
action with a planetesimal disk is important for the final
orbits of protoplanets. However, the disk mass of plan-
etesimals decreases due to the collisional cascade starting
from planetesimal fragmentation. Therefore, we need to
treat the evolution of orbits and masses consistently.
We apply the super-particle approximation for plan-
etesimals and smaller bodies ejected by collisional frag-
mentation; a super particle represents to planetesimals
and collisional fragments. Meanwhile, we apply a sin-
gle particle for a single protoplanet. We numerically
integrate the equations of motion of particles via the
fourth order Hermite scheme (Makino, & Aarseth 1992;
Kokubo, & Makino 2004). The orbital integration al-
lows us to accurately treat dynamical evolution and di-
rect collisions between protoplanets and super particles1.
However, the number of super particles that we apply is
much smaller than that of planetesimals and fragments
with which we are concerned so that statistical treat-
ment is required for accurate calculation of interactions
between super particles.
Morishima (2015) developed a method for the collisions
and dynamical interactions between super particles in N -
body simulations We treat the collisions between super
particles following Morishima (2015), while we ignore the
dynamical interaction because the collisional timescale is
much shorter than the dynamical interaction timescale
for dynamically hot planetesimals.
We consider the j-th super particle at the cylindrical
coordinate (rj, θj, zj) originated at the host star with
mass M∗. The surface density around the super particle,
Σj, is determined by the total mass of super particles in
the area with r = [rj − δrj : rj + δrj] and θ = [θj − δθj :
θj + δθj];
mj +
Nn,j(cid:88)
k
,
Σj =
1
4rjδrjδθj
mk
(1)
where Nn,j is the number of super particles in the area
and mk is the mass of the k-th super particle in the area.
We choose δr and δθ according to the accuracy of mass
evolution of planetesimals due to the collisional cascade,
which we discuss in §3.
The relative velocity, which characterises collisional
fragmentation of planetesimals in super particles, is de-
termined by the orbital elements of super particles in the
area. We calculate the relative velocity vj between plan-
etesimals in the j-th super particle with semimajor axis
aj, eccentricity ej, inclination ij, the longitude of peri-
center j, and the longitude of ascending node Ωj, given
by
where vK =(cid:112)GM∗/aj, G is the gravitational constant,
e2
r,j + i2
vr,j = vK
(cid:113)
r,j,
(2)
and
e2
r,j =
1
Nn,j
[e2
j + e2
k − 2ejek cos(j − k)],
(3)
1 The total ejecta mass caused by collisions between protoplan-
ets and super particles is estimated to be much smaller than the
disk mass and their collisional lifetimes are much shorter than that
of the disk. Therefore, we ignore collisional ejecta caused by colli-
sions between protoplanets and super particles.
Nn,j(cid:88)
k
Nn,j(cid:88)
k
i2
r,j =
1
Nn,j
[i2
j + i2
k − 2ijik cos(Ωj − Ωk)].
(4)
The collisional cascade grinds planetesimals down to
micron-sized grains, which are blown out due to the
radiation pressure of the host star (e.g., Kobayashi et
al. 2008, 2009; Krivov 2010). Collisional fragmenta-
tion of planetesimals thus reduces the surface density
of planetesimals due to the collisional cascade. We con-
sider the quasi-steady-state collisional cascade, for which
Kobayashi & Tanaka (2010) derived the reduction rate
of the surface density of planetesimals, given by
(cid:18) vr(mc)2
2QD(mc)∗
(cid:19)αc−1
f (αc),
(5)
dΣ
dt
= − (2 − αc)2
m1/3
c
Σ2ΩK
where mc is the mass of largest bodies in the collisional
cascade, αc is the index of the mass distribution of bod-
ies determined by the collisional cascade, f (αc) is the
dimensionless value dependent on αc, and Q∗
D is the spe-
cific impact energy needed for the ejection of the half
mass of colliders. The mass of largest bodies mc does
not change in the collisional cascade, which is valid until
D ∝ mp, αc is given
the bodies with mc exist. For v2
by (Kobayashi & Tanaka 2010)
r /Q∗
αc =
11 + 3p
6 + 3p
.
(6)
The mass distribution with the index αc is achieved in
the timescale for the collisional cascade. An earlier mass
distribution index may be different from αc in Eq. (6)
. However, the mass loss due to the collisional cascade
is negligible in such an early stage and the mass evolu-
tion mainly occurs when the mass distribution given by
Eq. (6) is achieved.
Therefore, Eqs. (5) and (6) are
valid to treat the surface density evolution accurately.
According to Kobayashi & Tanaka (2010), f (αc) is given
by
f (αc) = 1.055ρ−2/3
− ln +
(cid:20)(cid:18)
(cid:19)
1
2 − b
s1(αc) + s2(αc) + s3(αc)
(cid:21)
,
(7)
where ρ is the density of bodies, b is the power-law index
of the mass distribution of fragments, is the constant
determining the largest mass of fragments, and2
(cid:90) ∞
(cid:90) ∞
(cid:90) ∞
0
0
s1(αc) =
dφ
s2(αc) =−
φ1−αc
1 + φ
,
φ1−αc
1 + φ
dφ
φ−αc
1 + φ
(8)
(9)
ln
φ
(1 + φ)2 ,
0
dφ
s3(αc) =
ln(1 + φ).
(10)
The impact laboratory experiments shows b = 1.5 − 1.7
and ∼ 0.1 (e.g., Takagi et al. 1984; Nakamura, & Fu-
jiwara 1991) so that the choice of b and insignificantly
change the value of f (αc).3 Therefore, we set b = 3/5
2 We correct a typo in s3(αc) of Kobayashi & Tanaka (2010).
3 We here evaluate the dependence of f (αc) on b. We consider
p = 0.453, and then αc = 1.68. The values of s1, s2, and s3 are
calculated to be 3.7, 7.4, and 2.6, respectively. The function f (1.68)
with = 0.2 is estimated to be 28.9 and 33.8 for b = 1.5 and 1.7,
respectively.
Importance of Giant Impact Ejecta
3
and = 0.2.
If vr is fixed, the integration of Eq. (5) over time t gives
Σ(0)
,
1 + t/τ0
Σ(t) =
where τ0 = −Σ(0)/ Σ(0).
In the giant impact stage,
Eq. (11) is not always valid because of the evolution of
vr. Therefore, we use Eq. (11) only for the validation of
our simulation in §3.
(11)
Based on Eq. (5), the mass-loss rate of the j-th super
particle due to the collisional cascade is given by
(cid:32)
(cid:33)αc−1
1
mj
dmj
dt
= − (2 − αc)2
m1/3
c,j
ΣjΩK
v2
r,j
2QD(mc,j)∗
f (αc),
Figure 1. Relative errors of the surface density at t = 100 years
with the setting of τ0 ≈ 0.6 years for δθ = π/8 (red circles) and π
(blue squares), as a function of the number of super particles. The
error bars indicate the standard errors obtained from 40 runs.
(12)
where mc,j is the mass of the largest bodies in the j-th
super particle. We calculate the mass evolution of super
particles via the integration of Eq.
(12) using Eqs.(1)
and (2).
The masses of the largest bodies in super particles,
mc, are set to (cid:38) 1016 g corresponding to (cid:38) 1 km in ra-
dius. Therefore, Q∗
D of such a body is mainly determined
by shuttering and gravitational reaccumulation (Benz &
Asphaug 1999; Leinhardt, & Stewart 2012; Jutzi 2015;
Genda et al. 2015, 2017; Suetsugu et al. 2018). There-
fore, Q∗
D has a monotonous increasing function of mc,
which is assumed to be (e.g., Benz & Asphaug 1999)
(cid:18) mc
(cid:19)p
1021 g
Q∗
D = Q0
,
(13)
where Q0 and p are determined from numerical simula-
tions. The collisional cascade is controlled not by indi-
vidual collisions but by successive collisions. The value
averaged over impact angles is applied for Q∗
D. Tak-
ing into account self-gravity and a model of rock frac-
tures, Benz & Asphaug (1999) obtained the averaged
values for Q0 and p via the smoothed particle hydro-
dynamics (SPH) simulation. Recently, the dependence
of Q∗
D on the number of SPH particles is argued and
the value of Q∗
D decreases 20% in the limit of high res-
olution simulation (Genda et al. 2015). However, the
friction of damaged rock, which was not considered in
Benz & Asphaug (1999), increases Q∗
D (Jutzi 2015). The
value of Q∗
D obtained in Benz & Asphaug (1999) is simi-
lar to that given by the high-resolution simulations with
the fraction (Suetsugu et al. 2018). Therefore, we set
Q0 = 9.5 × 108 erg/g and p = 0.453 (Benz & Asphaug
1999).
3. VALIDATION FOR COLLISIONAL CASCADE
We perform a simulation for collisional evolution of a
planetesimal disk composed of 2,000 super particles with
a = 0.975 -- 1.025 AU and mc = 1016 g. The radial distri-
bution of super particles is put according to Σ(a) ∝ a−1
and e and i have the Rayleigh distributions with mode
values e = 0.01(a/1 AU)1/2 and i = 0.005(a/1 AU)1/2.
The collisional-cascade timescale in this setting is esti-
mated to be τ0 ≈ 0.6 years. We evaluate the accuracy
of the method from the comparison with the analytic
solution in Eq. (11) at t = 100 years.
The accuracy of the simulation depends on the number
of particles in the neighbor area, Nn. We calculate the
relative error of the surface density from comparison of
the simulation with the analytical solution. Fig. 3 shows
the dependence of relative errors on the number of su-
per particles. The errors are roughly proportional to the
number of super particles, which is proportional to Nn
for fixed δθ. However, the errors are almost independent
of δθ. Although Nn decreases with decreasing δθ, the
number of particles having the experience in a neighbor
area is almost same in the time much longer than orbital
periods if the number of super particles and δr are fixed.
The timescale of the collisional cascade is mainly much
longer than orbital periods. Therefore, the accuracy bet-
ter than 10% is obtained if the number of particles in the
annulus with width δr is (cid:38) 10.
4. ORBITAL EVOLUTION OF PROTOPLANETS IN A
SWARM OF FRAGMENTING PLANETESIMALS
We carry out simulations for the orbital evolution of
3 protoplanets with an Earth mass M⊕ in a planetesi-
mal disk with 30M⊕ composed of 3,000 super particles
around the central star with solar mass M(cid:12). The semi-
major axis of the intermediate protoplanet is initially set
at 1 AU, the orbital separation of protoplanets is 10 mu-
tual Hill radii, and their eccentricities and inclinations
are 0.03 and 0.015, respectively. The radial distribution
of super particles is initially put according to Σ(a) ∝ a−1
and their e and i have the Rayleigh distributions with
mode values e = 0.03 and i = 0.015, respectively. The in-
termediate radius and width of the planetesimal disk are
1 AU and 30 mutual Hill radii, respectively. For the treat-
ment of collisional fragmentation, we put δr = 0.01 AU
and δθ = π/8.
The collisional cascade is characterized by planetesi-
mal mass mc. We set mc = 1016 g (≈ 1 km in radius) for
collisional fragmentation, while we have an additional
D = ∞
simulation without collisional fragmentation (Q∗
or mc = ∞). Figure 4a,b shows the orbital distribution
of protoplanets and super particles at t = 103 years with
collisional fragmentation, while Figure 4c,d is that with-
out fragmentation. Dynamical friction decreases e and i
of protoplanets, while e and i of planetesimals increase.
Increases in e and i of planetesimals activate their colli-
sional cascade, which reduces the surface density of plan-
etesimals. Collisional fragmentation weakens dynamical
friction so that e and i of protoplanets with fragmenta-
tion remain higher than those without fragmentation.
Figures 4 -- 4 show the evolution of the root mean
squares of eccentricities (cid:104)e2(cid:105)1/2 and inclinations (cid:104)i2(cid:105)1/2
10-310-210-1100101102103 Relative ErrorNumber4
Figure 2.
Eccentricities (a,c) and inclinations in radian (b,d)
of super particles (red dots) and protoplanets (blue fulled circles)
with fragmentation (a,b) or without fragmentation (c,d).
for planetesimals and protoplanets in the same initial set-
ting of planetesimal disks and protoplanets as Figure 4.
If we ignore collisional fragmentation, (cid:104)e2(cid:105)1/2 and (cid:104)i2(cid:105)1/2
of planetesimals become much larger than those of proto-
planets (Figure 4). This is caused by dynamical friction.
However, collisional fragmentation of planetesimals de-
creases the surface density of planetesimals (Figures. 4
and 4), which suppresses the e and i reduction of proto-
planets via dynamical friction. Collisional fragmentation
is effective for small mc so that (cid:104)e2(cid:105)1/2 and (cid:104)i2(cid:105)1/2 of
protoplanets insignificantly change. On the other hand,
the e and i evolution for planetesimals are almost inde-
pendent of fragmentation. This is because the viscous
stirring of protoplanets controls e and i of planetesimals.
If we consider two populations of bodies such as pro-
toplanets with (cid:104)e2(cid:105)1/2
, and mass mp,1 and plan-
etesimals with (cid:104)e2(cid:105)1/2
, and mass mp,2, the time
differential of (cid:104)e2(cid:105)1/2
for α = 1 or 2 (proto-
planets or planetesimals) due to dynamical friction and
viscous stirring is analytically given by (Ohtsuki et al.
2002)
, (cid:104)i2(cid:105)1/2
, (cid:104)i2(cid:105)1/2
and (cid:104)i2(cid:105)1/2
α
α
2
2
1
1
(cid:88)
(cid:34)
(cid:34)
d(cid:104)e2(cid:105)α
dt
d(cid:104)i2(cid:105)α
dt
= a2
= a2
+
+
h4
α,βmp,β
Ns,β
0Ω
mp,β(cid:104)e2(cid:105)β − mp,α(cid:104)e2(cid:105)α
β=1,2
(mp,α + mp,β)2
(cid:88)
(cid:104)e2(cid:105)α + (cid:104)e2(cid:105)β
h4
α,βmβ
Ns,β
0Ω
mp,β(cid:104)i2(cid:105)β − mp,α(cid:104)i2(cid:105)α
β=1,2
(mp,α + mp,β)2
(cid:104)i2(cid:105)α + (cid:104)i2(cid:105)β
(cid:18)
(cid:19)(cid:21)
(cid:18)
(cid:19)(cid:21)
mp,βPVS
mp,βQVS
QDF
,
(15)
PDF
,
(14)
where a0 is the mean semimajor axis of bodies, Ns,β is
the surface number density of bodies for β = 1 or 2,
hα,β = [(mp,α + mp,β)/3M∗]1/3 is the reduced Hill radius
of bodies with masses mp,α and mp,β, PVS, PDF, QVS,
and QDF are the efficiencies for viscous stirring and dy-
namical friction for (cid:104)e2(cid:105)1,2 and (cid:104)i2(cid:105)1,2, respectively. The
analytical formulae of PVS, PDF, QVS, and QDF are given
as a function of (cid:104)e2(cid:105)1,2 and (cid:104)i2(cid:105)1,2 in Ohtsuki et al. (2002).
The e and i variation rates due to dynamical friction
are given by the second terms on the right hand sides
of Eqs.
(14) and (15), respectively. Protoplanets are
much more massive than planetesimals (mp,1 (cid:29) mp,2).
For mp,1(cid:104)e2(cid:105)1 (cid:29) mp,2(cid:104)e2(cid:105)2 and mp,1(cid:104)i2(cid:105)1 (cid:29) mp,2(cid:104)i2(cid:105)2,
the dynamical-friction damping rates for protoplanets is
approximated to be
(cid:18) mp,1
(cid:19)1/3
3M∗
,
d(cid:104)e2(cid:105)1
dt
1
PDF
∼ 1
QDF
d(cid:104)i2(cid:105)1
dt
∼ − a2
oΩΣ
6M∗
(16)
where Σ is the surface density of planetesimals and we
assume (cid:104)e2(cid:105)1 ≈ (cid:104)e2(cid:105)2 and (cid:104)i2(cid:105)1 ≈ (cid:104)e2(cid:105)2. The damping
rates are mainly determined by Σ but are almost inde-
pendent of the choice of planetesimal masses, mp,2. This
means the super-particle approximation is valid. How-
ever, dynamical friction leads to the equipartition of the
random energies such as mp,1(cid:104)e2(cid:105)1/2
and
mp,1(cid:104)i2(cid:105)1/2
. The values for the equiparti-
tion depend on the choice of mp,2. Therefore, we need
to care the choice of super-particle masses if we are in-
terested in the equipartition.
1 ∼ mp,2(cid:104)e2(cid:105)1/2
1 ∼ mp,2(cid:104)i2(cid:105)1/2
2
2
To compare the results of simulations with Eqs. (14)
and (15), we set mp,2 to be super-particle mass mj
and integrate Eqs. (14) and (15) over time (see Fig. 4).
The analytic solution is consistent with the simula-
tion. For t (cid:38) 100 years, dynamical friction is ineffec-
tive because of the achievement of energy equipartition
such as mp,1(cid:104)e2(cid:105)1/2
1 ∼
mp,2(cid:104)i2(cid:105)1/2
. The result depends on the super-particle
mass as discussed above. However, collisional fragmen-
tation mainly occur prior to the achievement of the
equipartition (see Figs. 4 and 4).
1 ∼ mp,2(cid:104)e2(cid:105)1/2
and mp,1(cid:104)i2(cid:105)1/2
2
2
The mass of planetesimals mp,2 in Eqs. (14) and (15)
is initially set to be super-particle masses mj, although
mc in Eq. (5) is independently given. We then integrate
Eqs. (5), (14), and (15) over time. The analytic solution
is roughly in agreement with the results of simulations
(see Figs. 4 and 4). In the simulations, each super parti-
cle has an independent mass. Super particles in the inner
disk effectively lose their masses via collisional fragmen-
tation because of high collisional speeds, while super par-
ticles tend to have large masses in the outer disk. On the
other hand, the mass evolution of planetesimal disks is
calculated with the averaged collisional speed in the an-
alytical solution. Therefore, the analytic solution cannot
perfectly reproduce the simulations. However, the ten-
dency of orbital interaction and collisional fragmentation
is understood from the analytical solution.
It should be noted that the relative increases in proto-
planet masses are smaller than 0.2 and 0.02 in the case
without and with collisional fragmentation, respectively.
Those are caused by collisions with planetesimals. We
estimate the effect of collisional damping due to the col-
lisional accretion according to the model by Kobayashi et
al. (2016), resulting in the e damping of 0.006 and 0.0001
for protoplanets without and with collisional fragmenta-
tion, respectively. The collisional damping insignificantly
affects the orbital evolution of protoplanets. Therefore,
the analytical solution without collisional damping is in
good agreement with the results of simulations.
The timescale of dynamical friction is estimated from
and
Eqs. (14) and (15).
2
(cid:104)e2(cid:105)1/2
damping
timescale due to dynamical friction has the relation as
for mp,1 (cid:29) mp,2, the (cid:104)e2(cid:105)1/2
1 (cid:29) mp,2(cid:104)e2(cid:105)1/2
2 (cid:29) (cid:104)e2(cid:105)1/2
If mp,1(cid:104)e2(cid:105)1/2
1
1
00.010.020.030.040.050.060.070.080.090.80.911.11.2(b)InclinationSemimajor Axis [AU]0.80.911.11.2(d)Semimajor Axis [AU]00.020.040.060.080.10.120.14(a)Eccentricity(c)Importance of Giant Impact Ejecta
Q0
9.5 × 108erg/g
(cid:19)0.68
(18)
5
This estimate implies a planetesimal disk with Mdisk (cid:28)
M⊕ leads to eccentricity damping for Earth-sized proto-
planets in a long timescale (cid:29) 104yr via dynamical fric-
tion. In addition, dynamical friction increases (cid:104)e2(cid:105)1/2 of
planetesimals. The increase in (cid:104)e2(cid:105)1/2
2 makes τdf longer.
On the other hand, the decreasing timescale for Mdisk
is estimated from
due to the collisional cascade, τcc,
Eq. (5) to be (Kobayashi & Tanaka 2010)
(cid:18) Mdisk
(cid:19)0.79(cid:18) ∆a
(cid:19)−1(cid:32)(cid:104)e2(cid:105)1/2
(cid:19)(cid:16) a0
(cid:33)−1.36(cid:18)
(cid:17)4.18
2
0.1
M⊕
1022 g
0.1a0
1 AU
yr.
τcc ≈ 4 × 103
(cid:18) mc
×
1
2
during the (cid:104)e2(cid:105)1/2
For 100 km-sized planetesimals (mc ≈ 1022 g) with
(cid:104)e2(cid:105)1/2
2 ∼ 0.1, τcc (cid:46) τdf . It should be noted that dynami-
cal friction increases (cid:104)e2(cid:105)1/2
damping,
which shortens τcc and elongates τdf as discussed above.
Therefore, even if τcc ∼ τdf initially, τcc eventually be-
comes much shorter than τdf .
Equations (17) and (18) show τdf /τcc is independent of
Mdisk and ∆a but depends on (cid:104)e2(cid:105)1/2
, and mc.
For (cid:104)e2(cid:105)1/2
2 = 0.03, PDF ≈ 60 so that τcc (cid:28) τdf
for mc = 1016 g, while τdf ∼ τcc for mc ∼ 1019 g.
Therefore, as shown in Figs. 4 and 4, e and i of proto-
planets insignificantly decrease for mc = 1016 g because
of τcc (cid:28) τdf , while those are moderately damped for
τcc ∼ τdf with mc = 1019 g. Therefore, the condition
with tcc (cid:29) τdf is required for the e damping of proto-
planets.
1 = (cid:104)e2(cid:105)1/2
, (cid:104)e2(cid:105)1/2
1
2
Planets formed via giant impacts have high eccentric-
ities and inclinations. For Earth-sized planets formed
in the giant impact stages, the mean eccentricity is
∼ 0.1 (Chambers 2001; Kokubo et al. 2006), which is
much larger than the current eccentricities of Earth and
Venus. The depletion timescale of a planetesimal disk
with mc (cid:46) 1021 g is shorter than the e damping timescale
via dynamical friction (see Eqs. 17 and 18). Therefore,
larger planetesimals are required for e damping of Earth-
sized planets. We consider such large planetesimals are
produced from a giant impact. The mass ratios of largest
giant impact ejecta to parent protoplanets are ∼ 0.01
(Genda et al. 2015). Therefore, we consider mc ∼ 1026 g,
resulting in τcc (cid:29) τdf .
We perform the simulation for the orbital evolution of
a high-eccentricity planet with mass M⊕ in a swarm of gi-
ant impact ejecta. Giant impact ejecta initially have sim-
ilar orbits to parent planets. However, and Ω are even-
tually distributed uniformly due to perturbation from
other planets. The timescale to achieve a uniform distri-
bution for and Ω, which is roughly given by the preces-
sion rates for and Ω obtained from the secular pertur-
bation theory (Murray, & Dermott 1999), is estimated
as ∼ 105 − 106 years for giant impact ejecta around 1AU
perturbed by a Venus-like planet, which is much shorter
than the timescale for dynamical friction caused by giant
impact ejecta. Therefore, instead of ignoring perturba-
tion from other planets, we uniformly set and Ω of
giant impact ejecta from the beginning. We initially set
a = 1 AU and e = 2i = 0.1 for the planet and a = 0.95 --
Figure 3. The root mean squares of eccentricities (cid:104)e2(cid:105)1/2 (red)
and inclinations (cid:104)i2(cid:105)1/2 (blue) for planetesimals (open circles) and
protoplanets (filled circles) without collisional fragmentation in
the planetesimal disk same as Fig. 4. The error bars are given
by the standard deviation of 13 runs. Analytic solutions (cid:104)e2(cid:105)1/2
and (cid:104)i2(cid:105)1/2 for planetesimals (dotted lines) and protoplanets (solid
lines) is obtained from the time integration of Eqs. (14) and (15).
Figure 4. Same as Fig. 4 but for the treatment of collisional
fragmentation. We take into account collisional fragmentation for
mc = 1 × 1016 g. The mean surface density of planetesimals, Σ,
decreases due to collisional fragmentation. The ratio of Σ to initial
surface density Σ0 obtained from the simulation (open circles) and
the analytical solution (solid line) is shown in the bottle panel.
Figure 5. Same as Fig. 4 but for mc = 1 × 1019 g.
(Ohtsuki et al. 2002)
(cid:18) Mdisk
(cid:19)(cid:16) a0
(cid:19)−1(cid:32)(cid:104)e2(cid:105)1/2
(cid:17)3/2(cid:18) M∗
2
0.1
M⊕
(cid:33)2(cid:18) PDF
(cid:19)1/2
17
(cid:19)−1
10rH
1 AU
M(cid:12)
yr,
(17)
τdf ≈ 1 × 104
(cid:18) ∆a
×
where Mdisk is the total mass of a planetesimal disk, ∆a
is the width of the disk, and rH = (mp,1/3M∗)1/3a0 is the
Hill radius for protoplanets. Note that PDF is estimated
to be ∼ 10 for mp,1 ∼ M⊕ and (cid:104)e2(cid:105)1/2
2 ∼ 0.1.
1 ≈ (cid:104)e2(cid:105)1/2
0.0030.010.030.10.11101001000< e 2 >1/2, < i 2 >1/2Time [year]0.0030.010.030.1< e 2 >1/2, < i 2 >1/210-410-310-210-11000.11101001000Σ / Σ0 Time [year]0.0030.010.030.1< e 2 >1/2, < i 2 >1/210-410-310-210-11000.11101001000Σ / Σ0 Time [year]6
1.05 AU and (cid:104)e2(cid:105)1/2 = 2(cid:104)i2(cid:105)1/2 = 0.1 for a giant-impact-
ejecta disk composed of 120 super particles with total
mass 0.2M⊕ and mc = 1026 g (Fig. 4). The evolutions of
(cid:104)e2(cid:105)1/2 and (cid:104)i2(cid:105)1/2 for giant impact ejecta and the planet
differs from those predicted by the analytical solutions
given by the time integration of Eqs. (5), (14), and (15).
Once (cid:104)e2(cid:105)1/2 and/or (cid:104)i2(cid:105)1/2 are much larger than 0.1,
their orbits are controlled by the higher order terms for
(cid:104)e2(cid:105)1/2 and (cid:104)i2(cid:105)1/2, which are ignored in Eqs.(14) and
(15). The higher order terms make dynamical friction
less effective. Therefore, the variations of (cid:104)e2(cid:105)1/2 and
(cid:104)i2(cid:105)1/2 for the planet and planetesimals are less than
those expected by the analytic solution. However, e of
the planet becomes comparable to or larger than the an-
alytic estimate until 10 Myrs and decreases much greater
than the analytic estimate in several 10 Myrs. This is
caused by the orbital energy damping for the planet
rather than its angular momentum variation.
If the
angular momentum is fixed, the energy damping given
by −∆a/a results in ∆e = (1 − e2)∆a/2ae, where ∆e
and ∆a are the changes in e and a, respectively. There-
fore, even a small energy damping of ∆a/a ≈ −0.02 leads
to ∆e ≈ −0.1 for the planet. Giant impact ejecta are
scattered by the planet and tend to stay in the outer
disk so that the planet loses the orbital energy via the
scatterings. This is clearly shown in the semimajor-axis
evolution for the planet and giant impact ejecta (Fig. 4).
The collisional damping between the planet and giant im-
pact ejecta also reduces the orbital energy of the planet.
The accretion mass of giant impact ejecta onto the planet
is 0.086+0.004
−0.005M⊕ for t = 10 --
30 Myrs, and 0.029+0.04−0.004M⊕ for t = 30 -- 100 Myrs. Al-
though the e damping is significant after 30 Myrs, the
collisional damping expected from the accretion mass is
slight. Therefore, the energy loss for the planet due to
the scattering of giant impact ejecta mainly decreases its
eccentricity.
−0.006M⊕ in 10Myrs, 0.029+0.011
The mass evolution of giant impact ejecta due to col-
lisional fragmentation is less important than that esti-
mated by Eq. (5). Increase in e of giant impact ejecta
due to planetary perturbation makes the giant-impact-
ejecta disk wider. This decreases the surface density of
giant impact ejecta, which reduces the efficiency of col-
lisional fragmentation.
In addition, collisions between
high-eccentricity giant impact ejecta are more frequent
around the planetary orbit. Collisional fragmentation
between giant impact ejecta with similar and Ω thus
mainly occurs so that those relative velocities are com-
paratively small (see Eqs. 3 and 4). The mass loss by col-
lisional fragmentation is thus less effective due to the in-
crease of ejecta eccentricity. Therefore, the giant-impact-
ejecta disk is maintained in a long timescale (cid:38) 10 Myr,
resulting in the significant eccentricity damping for the
planet.
5. DISCUSSION
In the giant impact stage, orbital instability of proto-
planets occurs in a timescale longer than 10 Myrs (Cham-
bers et al. 1996; Chambers, & Wetherill 1998; Iwasaki et
al. 2001, 2002; Kominami, & Ida 2002; Kokubo et al.
2006). The protoplanets formed from the long-term evo-
lution mainly have high eccentricities (∼ 0.1) (Chambers
2001; Kokubo et al. 2006), which are much larger than
Figure 6. The orbital and mass evolution of a planet and a swarm
of giant impact ejecta. Top panel shows (cid:104)e2(cid:105)1/2 (red) and (cid:104)i2(cid:105)1/2
(blue) for the protoplanet (filled circles) and giant impact ejecta
(open circles). Middle panel also shows (cid:104)a(cid:105) (green). Bottom panel
represents the total mass of giant impact ejecta. We carry out
three runs with different initial positions. The error bars indicate
the minimum and maximum values. The lines are the same as
Fig. 4.
the current eccentricities of Earth and Venus. A swarm
of planetesimals may damp the eccentricities of proto-
planets due to dynamical friction (O'Brien et al. 2006;
Raymond et al. 2009; Morishima et al. 2010). Instead,
dynamical friction increases eccentricities of planetesi-
mals, which induces collisional fragmentation between
planetesimals. The collisional cascade grinds planetes-
imals until small fragments are brown out by radiation
pressure, which results in the mass loss of the planetes-
imal disk. Collisional fragmentation may thus suppress
the eccentricity damping for protoplanets. Therefore, we
have investigated the orbital interaction between proto-
planets and planetesimals, taking into account collisional
fragmentation.
Eccentricities and inclinations of protoplanets are
damped via dynamical friction if we ignore collisional
fragmentation (see Fig. 4cd). However, collisional frag-
mentation weakens dynamical friction (see Fig. 4ab).
The mass loss due to collisional fragmentation depends
on the typical planetesimal size of which planetesimals
mainly determine the total mass of the planetesimal disk.
Collisional depletion occurs in a short timescale for plan-
etesimal disks with small typical planetesimal sizes. For
100km-sized or smaller planetesimals, the eccentricity
damping is ineffective due to collisional fragmentation
(Figs. 4 and 4).
The primordial planetesimals may remain even in the
giant impact stage. Although the typical size of primor-
dial planetesimals is not unknown, the size may be on the
order of 100 km, similar to that of Main-Belt asteroids
(e.g., Kobayashi et al. 2016). Due to collisional frag-
mentation, dynamical friction for eccentricity damping
is ineffective in such a planetesimal disk. On the other
hand, the increase in the eccentricities of protoplanets is
required for the onset of the orbital instability of proto-
planets for giant impacts (Iwasaki et al. 2001). Although
0.010.030.10.3< e 2 >1/2, < i 2 >1/211.11.21.31.41.51.6< a > 0.030.10.3105106107108Mass [M⊕] Time [year]Importance of Giant Impact Ejecta
7
primordial planetesimal disks may not explain the small
eccentricities of Earth and Venus, such a disk does not
inhibit the onset and maintain of giant impact stages
(Walsh, & Levison 2019). In addition, the direct forma-
tion of Earth and Venus via the accretion of planetesi-
mals is insignificant due to the depletion of remnant plan-
etesimal disks via collisional fragmentation (Kobayashi et
al. 2010; Kobayashi & Dauphas 2013). Therefore, colli-
sional fragmentation supports the giant impact scenario
to form Venus and Earth.
Giant impacts lead to the ejection of fragments as well
as collisional growth of protoplanets. The collisional
ejecta from single impacts have 0.1 -- 0.3M⊕, while the
typical size of giant impact ejecta can be as large as
1,000 km (Genda et al. 2015). The simulation for the in-
teraction between a protoplanet and giant impact ejecta
shows significant eccentricity damping of the protoplanet
in (cid:38) 30 Myr (see Fig. 4). The collisional fragmentation is
less effective in an originally narrow planetesimal disk for
high-eccentricity planetesimals, while the dynamical fric-
tion is also less effective for high eccentricities and incli-
nations of planetesimals. As a result, eccentricity damp-
ing for planets occurs in a long timescale ∼ 100 Myr.
Giant impact ejecta are produced even in the early
giant impact stage (Genda et al. 2015). However, the
orbital separations of protoplanets are so narrow that gi-
ant impact ejecta are distributed widely due to interac-
tions with other protoplanets. In addition, the eccentric-
ity damping timescale due to dynamical friction is much
longer than collisional timescale between protoplanets.
Therefore, sequent giant impacts occurs due to insignif-
icance of dynamical friction by giant impact ejecta. On
the other hand, stable orbital configurations of proto-
planets are achieved after tens of giant impacts. Such
orbital separations are wide enough to keep the orbital
concentration of giant impact ejecta around the orbits of
single protoplanets. Therefore, dynamical friction by gi-
ant impact fragments is effective in the late giant impact
stage, which may form low-eccentricity planets.
Finally we need to discuss the accuracy of the mass loss
due to collisional fragmentation. As we discuss above,
the collisional mass loss mainly occurs after the colli-
sional cascade is achieved. In the collisional cascade, the
mass loss is mainly determined by the total ejecta mass
from single collisions, and is insensitive to the mass dis-
tribution of ejecta (Kobayashi & Tanaka 2010). There-
fore, the uncertainty of collisional fragmentation mainly
comes from Q∗
D at the typical sizes of planetesimals (see
Eq. 18). As discussed in the previous studies (Kobayashi
et al. 2016; Kobayashi, & Tanaka 2018), Q∗
D of 10 km-
sized or smaller primordial planetesimals may be much
larger than Q∗
D we set in the simulations. However, tak-
ing into account the enhancement of Q∗
D, 100 km-sized
or smaller primordial planetesimals insignificantly work
for the e damping of protoplanets. On the other hand,
Q∗
D of 1000 km-sized bodies are mainly determined by
the self-gravity of colliding bodies (Kobayashi et al. 2010,
2011; Genda et al. 2015, 2017; Suetsugu et al. 2018). The
uncertainty is small for Q∗
D of largest ejecta of giant im-
pacts. Therefore, the e damping of planets formed in the
giant impact stage is likely to be caused by giant impact
ejecta.
6. SUMMARY
Terrestrial planets are formed via giant impacts be-
tween Mars-sized protoplanets. The resultant planets
have larger eccentricities than the current values for
Earth and Venus. Dynamical friction with a planetes-
imal disk is expected to damp the eccentricities of pro-
toplanets. On the other hand, the collisional cascade
of planetesimals and blow-out of small collisional frag-
ments by radiation pressure decrease the planetesimal-
disk mass, which weakens dynamical friction with plan-
etesimals. Therefore, we have investigated the orbital
evolution of planets with collisional fragmentation. Our
findings are as follows.
1. We have developed an N-body simulation code in-
volving the mass loss due to the collisional cascade.
We have calculated the mass evolution of a plan-
etesimal disk composed of super particles using the
code, which reproduces the analytical solution for
the mass loss due to the collisional cascade with
high accuracy.
2. We have investigated the evolution of orbits and
masses of protoplanets and planetesimals via grav-
itational interaction and collisional fragmentation
using the N-body code. If collisional fragmentation
is ignored, dynamical friction damps eccentricities
and inclinations for protoplanets. However, colli-
sional fragmentation suppresses dynamical friction.
The timescale ratio of dynamical friction to colli-
sional fragmentation depends on the typical plan-
etesimal size but not on the disk mass. Even a mas-
sive planetesimal disk composed of 100km-sized or
smaller planetesimals cannot damp eccentricities of
planets. Therefore, the disks composed of primor-
dial planetesimals are ineffective for the eccentric-
ity damping for planets.
3. Giant impacts eject collisional fragments. The to-
tal masses of giant impact ejecta are several tenth
of colliding planets. The typical size of the largest
ejecta is ∼ 1000 km. We have carried out sim-
ulations for a planet with an Earth mass in the
disk with 0.2 Earth masses composed of giant im-
pact ejecta with largest ejecta 1026 g. Collisional
mass loss is insignificant for such a large typical
size. Eccentricity of the planet is damped from
0.1 to ∼ 0.01 due to interactions with giant impact
ejecta in (cid:38) 30 Myrs. Therefore, giant impact ejecta
are possible to decrease the eccentricities of planets
formed via giant impacts, comparable to those of
Earth and Venus.
We thank the reviewer for beneficial comments. The
work is supported by Grants-in-Aid for Scientific Re-
search (17K05632, 17H01105, 17H01103, 18H05436,
18H05438) from MEXT of Japan and by JSPS Core-to-
Core Program "International Network of Planetary Sci-
ences".
REFERENCES
Benz, W., & Asphaug, E. 1999,Icarus, 142, 5
Chambers, J. E. 2001, Icarus, 152, 205
Chambers, J. E., & Wetherill, G. W. 1998, Icarus, 136, 304
Chambers, J. E., Wetherill, G. W., & Boss, A. P. 1996, Icarus,
119, 261
8
Genda, H., Kobayashi, H., & Kokubo, E. 2015, ApJ, 810, 136
Genda, H., Fujita, T., Kobayashi, H., et al. 2015, Icarus, 262, 58
Genda, H., Fujita, T., Kobayashi, H., et al. 2017, Icarus, 294, 234
Iwasaki, K., Emori, H., Nakazawa, K., et al. 2002, PASJ, 54, 471
Iwasaki, K., Tanaka, H., Nakazawa, K., et al. 2001, PASJ, 53, 321
Jutzi, M. 2015, Planet. Space Sci., 107, 3
Kobayashi, H., & Dauphas, N. 2013, Icarus, 225, 122
Kobayashi, H., & Lohne, T. 2014, MNRAS, 442, 3266
Kobayashi, H., & Tanaka, H. 2010, Icarus, 206, 735
Kobayashi, H., Tanaka, H., Krivov, A. V., & Inaba, S. 2010,
Icarus, 209, 836
Kobayashi, H., Tanaka, H., & Krivov, A. V. 2011, ApJ, 738, 35
Kobayashi, H., & Tanaka, H. 2018, ApJ, 862, 127
Kobayashi, H., Tanaka, H., & Okuzumi, S. 2016, ApJ, 817, 105
Kobayashi, H., Watanabe, S.-. ichiro ., Kimura, H., et al. 2008,
Icarus, 195, 871
Levison, H. F., Duncan, M. J., & Thommes, E. 2012, AJ, 144, 119
Leinhardt, Z. M., & Stewart, S. T. 2012, ApJ, 745, 79
Makino, J., & Aarseth, S. J. 1992, PASJ, 44, 141
Morishima, R. 2015, Icarus, 260, 368
Morishima, R., Stadel, J., & Moore, B. 2010, Icarus, 207, 517
Morishima, R. 2017, Icarus, 281, 459
Murray, C. D., & Dermott, S. F. 1999, Solar system dynamics by
C.D. Murray and S.F. McDermott. (Cambridge
Nakamura, A., & Fujiwara, A. 1991, Icarus, 92, 132
Nesvold, E. R., Kuchner, M. J., Rein, H., et al. 2013, ApJ, 777,
144
O'Brien, D. P., Morbidelli, A., & Levison, H. F. 2006, Icarus, 184,
39
Ohtsuki, K., Stewart, G. R., & Ida, S. 2002, Icarus, 155, 436
Raymond, S. N., O'Brien, D. P., Morbidelli, A., et al. 2009,
Icarus, 203, 644
Kobayashi, H., Watanabe, S.-. ichiro ., Kimura, H., et al. 2009,
Suetsugu, R., Tanaka, H., Kobayashi, H., et al. 2018, Icarus, 314,
Icarus, 201, 395
Kokubo, E., & Makino, J. 2004, PASJ, 56, 861
Kokubo, E., Kominami, J., & Ida, S. 2006, ApJ, 642, 1131
Kominami, J., & Ida, S. 2002, Icarus, 157, 43
Kral, Q., Th´ebault, P., & Charnoz, S. 2013, A&A, 558, A121
Krivov, A. V. 2010, Research in Astronomy and Astrophysics, 10,
383
121
Takagi, Y., Mizutani, H., & Kawakami, S.-I. 1984, Icarus, 59, 462
Walsh, K. J., & Levison, H. F. 2019, Icarus, 329, 88
|
1705.06090 | 1 | 1705 | 2017-05-17T11:08:08 | Observational evidence for two distinct giant planet populations | [
"astro-ph.EP"
] | Analysis of the statistical properties of exoplanets, together with those of their host stars, are providing a unique view into the process of planet formation and evolution. In this paper we explore the properties of the mass distribution of giant planet companions to solar-type stars, in a quest for clues about their formation process. With this goal in mind we studied, with the help of standard statistical tests, the mass distribution of giant planets using data from the exoplanet.eu catalog and the SWEET-Cat database of stellar parameters for stars with planets. We show that the mass distribution of giant planet companions is likely to present more than one population with a change in regime around 4\,M$_{Jup}$. Above this value host stars tend to be more metal poor and more massive and have [Fe/H] distributions that are statistically similar to those observed in field stars of similar mass. On the other hand, stars that host planets below this limit show the well-known metallicity-giant planet frequency correlation. We discuss these results in light of various planet formation models and explore the implications they may have on our understanding of the formation of giant planets. In particular, we discuss the possibility that the existence of two separate populations of giant planets indicates that two different processes of formation are at play. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. santos
May 18, 2017
c(cid:13)ESO 2017
Observational evidence for two distinct giant planet populations
N. C. Santos1, 2, V. Adibekyan1, P. Figueira1, D. T. Andreasen1, 2, S. C. C. Barros1, E. Delgado-Mena1, O.
Demangeon1, J. P. Faria1, 2, M. Oshagh3, S. G. Sousa1, P. T. P. Viana1, 2, and A.C.S. Ferreira1, 2
1 Instituto de Astrofísica e Ciências do Espaço, Universidade do Porto, CAUP, Rua das Estrelas, 4150-762 Porto, Portugal
2 Departamento de Física e Astronomia, Faculdade de Ciências, Universidade do Porto, Rua do Campo Alegre, 4169-007 Porto,
Portugal
3 Institut für Astrophysik, Georg-August-Universität, Friedrich-Hund-Platz 1, 37077 Göttingen, Germany
7
1
0
2
y
a
M
7
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
0
9
0
6
0
.
5
0
7
1
:
v
i
X
r
a
Received date / Accepted date
ABSTRACT
Context. Analysis of the statistical properties of exoplanets, together with those of their host stars, are providing a unique view into
the process of planet formation and evolution.
Aims. In this paper we explore the properties of the mass distribution of giant planet companions to solar-type stars, in a quest for
clues about their formation process.
Methods. With this goal in mind we studied, with the help of standard statistical tests, the mass distribution of giant planets using
data from the exoplanet.eu catalog and the SWEET-Cat database of stellar parameters for stars with planets.
Results. We show that the mass distribution of giant planet companions is likely to present more than one population with a change
in regime around 4 MJup. Above this value host stars tend to be more metal poor and more massive and have [Fe/H] distributions that
are statistically similar to those observed in field stars of similar mass. On the other hand, stars that host planets below this limit show
the well-known metallicity-giant planet frequency correlation.
Conclusions. We discuss these results in light of various planet formation models and explore the implications they may have on our
understanding of the formation of giant planets. In particular, we discuss the possibility that the existence of two separate populations
of giant planets indicates that two different processes of formation are at play.
Key words. Planetary systems; Techniques: spectroscopic; Stars: abundances; Methods: statistical; Planets and satellites: formation
-----------------
1. Introduction
The detection of more than 3500 planets orbiting solar-type stars
(e.g. exoplanet.eu – Schneider et al. 2011) makes the exoplanet
domain prone to statistical studies. This opens the possibility to
analyse the properties of newly found worlds and, in comparison
with the model predictions, better understand the processes of
planet formation and evolution (see e.g. Mayor et al. 2014).
In this context, a significant amount of information was
brought by the analysis of the star-planet connection.
Initial
findings have shown that the planet occurrence is closely linked
to the metallicity of the host star (Gonzalez 1997; Santos et al.
2001, 2004; Ida & Lin 2004; Fischer & Valenti 2005; Mordasini
et al. 2012a), even if this trend is a matter of debate for planets
hosted by giant stars (Pasquini et al. 2007; Reffert et al. 2015).
Stellar mass has also been suggested to influence severely the
planet formation efficiency (Johnson et al. 2007; Lovis & Mayor
2007; Bonfils et al. 2013; Kennedy & Kenyon 2008; Reffert et al.
2015). Furthermore, several studies have shown that the archi-
tecture of the planetary systems is closely connected to the prop-
erties of the star. For example, it is now known that the orbital
period and eccentricity of the planets may depend on the chem-
ical content of the star (Dawson & Murray-Clay 2013; Beaugé
& Nesvorný 2013; Adibekyan et al. 2013). In brief, the under-
standing of planet formation and evolution processes is tightly
connected to the understanding of the properties of their host
stars.
In this paper we explore the mass distribution of giant plan-
ets orbiting solar-type stars in a search for clues about the for-
mation of giant planets. In Sect. 2 we present evidence that this
distribution likely presents more than one population, with two
regimes separated in planet mass. In Sect. 3 we then explore the
properties of the host stars of two planet regimes, to show that
a significant difference exists concerning the stellar metallicity
and stellar mass. In Sect. 4 the results are discussed in light of
the different models of planet formation and evolution.
2. The mass distribution of giant planets
In Fig. 1 we plot the mass distribution for all giant planets with
masses between 1 and 20 MJup orbiting solar-type stars as listed
in the exoplanet.eu database (Schneider et al. 2011)1. The lower
limit was set because we only want to explore the giant planet
domain. We only selected planets discovered through the transit
and radial velocity (RV) methods2.
In the upper panel of Fig. 1 we plot the mass distribution in
log scale, while in the lower panel we present the same distri-
bution in linear scale. The upper panel suggests that the mass
distribution may have two different maxima, separated by a val-
ley at around 4 MJup. These two regimes are also denoted in
the linear scale plot: below ∼4 MJup the distribution has a clear
1 Data from the 16th of February 2017.
2 These constitute the vast majority of the known planets with a mass
estimate. In the case of planets only detected with radial velocities, the
masses represent minimum masses.
Article number, page 1 of 6
stars using the following criteria. First, we selected only stars
for which stellar temperatures, log g, metallicities, and masses
are listed in SWEET-Cat and that have visual magnitudes lower
than 13. This magnitude limit, though not very constraining,
ensures that the planet masses can be derived with a reason-
able confidence using radial velocities, and that the stellar pa-
rameters, namely effective temperature and metallicity, could be
derived with a high level of reliability based on high quality
spectroscopic data. Then, we selected only stars with temper-
atures below 6500 K and above 4000 K, thus excluding targets
for which stellar atmospheric parameters could be less reliable.
We also considered only planets with masses below 15 MJup, to
try to avoid the inclusion of brown-dwarf companions. Finally,
we decided to cut in orbital period to select only planets with
periods longer than 10 days and shorter than 5 years. The first
limit allows us to avoid hot Jupiters, whose formation and migra-
tion processes are largely debated (e.g. Ngo et al. 2016; Nelson
et al. 2017). The upper limit allows us to guarantee that the sam-
ple is reasonably complete; giant planets in longer period orbits
could have been missed in radial velocity surveys. However, we
tested the results without including any constraints on the shorter
period limit and adding a stronger constraint on the upper value
(e.g. up to 2.5 years)4. The conclusions of the analysis presented
below become even stronger when we include hot Jupiters or if
we only select planets up to periods of just 2.5 years. A table
with the selected sample is available in electronic form.
3.1. Planet mass, stellar mass, and stellar metallicity
Using this selected sample, in Fig. 2 we compare the [Fe/H] dis-
tribution of the stars with planet masses above and below 4 MJup.
In the plots we use a "normalized number" to better compare
the various populations (each with a different number of stars),
meaning that the histograms are normalized such that the inte-
gral over the [Fe/H] range is 1. The upper left panel presents this
distribution for all the hosts, while the remaining panels present
the distributions for stars within different mass intervals. Stellar
masses were taken from SWEET-Cat, and were derived using
the calibration in Torres et al. (2010)5.
For each case, we applied a Kolmogorov-Smirnov (K-S) test
to explore whether the two samples (more and less massive plan-
ets) come from the same parent population6. The results of this
analysis are presented in Table 1, and show that if we use all the
stars or only the stars with mass above 1.5 M(cid:12) the low p-values
strongly suggest the two samples are not part of the same popu-
lation. On the other hand, as we go towards lower stellar mass
regimes, the K-S p-values increase, i.e., there a higher evidence
for one single population (see also Adibekyan et al. 2013). How-
ever, in all stellar mass regimes the hosts of the more massive
planets always have lower average metallicity values. Also, the
spread of the metallicity distribution of the stars with the most
massive planets, as measured via the standard deviation (STD),
4 A 2 MJup planet in a five-year period circular orbit around a solar-
mass star induces a semi-amplitude signal with 11 m s−1, which is a
value that is within the detection capabilities of present day instrumen-
tation; this value, however, it is not straightforward to detect if the planet
is orbiting noisier, giant stars.
5 The obtained results do not change if we use the masses listed in
exoplanet.eu.
6 Using the python scipy.stats.ks_2samp library.
ilar
(scipy.mstats.ks_2samp)
or
(scipy.stats.anderson_ksamp).
Sim-
a one-sided K-S test
test
Anderson-Darling
a
results
are obtained if we use
Fig. 1. Mass distribution of giant planets around solar-type stars, in
log (top panel) and linear (lower panel) scales. The green histograms
include all planets in exoplanet.eu, and the blue histogram includes only
planets following the criteria mentioned in Section. 3.
peaked shape, while above this mass the distribution is suddenly
mostly flat, with a slowly decreasing trend up to (tentatively)
∼15 MJup. The two dashed lines in the plot denote this change
in slope, and are used here merely as visual guide. The existence
of a transition at a similar mass value (∼5 MJup) has also been
recently mentioned by Bashi et al. (2017) based on the analysis
of planets with measured mass and radius, even if these authors
concentrated their discussion exclusively on planets with masses
lower than this limit.
The possible existence of two mass regimes suggests that
there may be two different populations of giant planets, with
masses above and below ∼4 MJup.
3. Exploring the various mass regimes
To explore this possibility, we compare the metallicity and mass
distributions for the stars having planets in these two planet-mass
regimes. For that purpose we started by compiling stellar pa-
rameters for all the stars selected above from the SWEET-Cat
database3 dabatase (Santos et al. 2013), which is a large cata-
log of stellar parameters for stars with planets. We then selected
3 https://www.astro.up.pt/resources/sweet-cat
Article number, page 2 of 6
N. C. Santos et al.: Observational evidence for two distinct giant planet populations
is always larger than that found for the lower mass planet host
stars. In other words, they span a larger metallicity range.
We also ran a clustering analysis on the two-dimensional
planet mass - stellar metallicity distribution using the planets se-
lected following the criteria mentioned in the next section. The
upper panel of Fig.3 shows a scatter plot of the above-mentioned
variables. The plot shows a clustering for planet masses below
∼4 MJup and metallicities above ∼−0.3 dex. Planets with masses
above that value are more scattered in both planet mass and stel-
lar metallicity. In the lower panel we show the result of a cluster-
ing analysis. A Gaussian mixture was considered7, in which the
data set is modelled with a fixed number of (two-dimensional)
Gaussian distributions. We assumed the presence of two clus-
ters and iteratively optimized the parameters of the distributions
using the expectation-maximization algorithm (Gupta & Chen
2011). In the lower panel of Fig.3 we show the underlying dis-
tribution of our sample and the two clusters that result from this
analysis. The clusters basically divide the data set into lower and
higher mass planets and overlap near 4 MJup. Their centres show
an offset in metallicity of 0.14 dex, where the average [Fe/H] is
lower for the hosts of the more massive planets.
It is worth noting that the higher mass stars in our sample are
also, on average, more evolved. The evolved stars (defined here
as those with log g<3.5) in our sample have mass values from
∼1 to 4.5 M(cid:12), while the dwarfs have maximum masses of about
1.5 M(cid:12) with most stars in the range between 0.7 and 1.3 M(cid:12).
Also, above 3 M(cid:12) all eight stars with planets in the mass range
above 4 MJup have metallicities below solar ([Fe/H] between -
0.13 and -0.74 dex), while the only star with a lower mass planet
has super-solar metallicities (0.07 dex).
3.2. Comparing with the solar neighbourhood
It is interesting to see that for the most massive stars and for those
with mass below solar mass, the average metallicity of the mas-
sive planet hosts is similar to the average metallicity of the so-
lar neighbourhood for dwarfs (−0.11 dex, STD=0.24) and giants
(−0.08 dex, STD=0.18) (Sousa et al. 2011; Alves et al. 2015). In
other words, in these stellar mass regimes, the metallicity distri-
bution of the stars with the most massive planets is very similar
to that observed in solar neighbourhood stars. Stars in the inter-
mediate mass range, on the other hand, present higher metallicity
values, on average. This likely reflects the fact that most of these
are dwarfs; as shown in Santos et al. (e.g. 2013), when a cut in
colour is applied to a sample of dwarfs, more massive stars are
on average more metal rich.
To explore these points in more detail, we compare the
metallicity distributions of the hosts of massive planets with
those of field dwarfs and giants, dividing the stars in the three
groups of stellar mass as listed in Table 1. The comparison of
the [Fe/H] distribution of the stars with M(cid:63) ≤1.0 M(cid:12) shows that
the two distributions have a K-S p-value of 0.04. Field dwarfs
(from Sousa et al. 2011) have slightly lower average metallic-
ity and standard deviation. Considering dwarfs with masses be-
tween 1.0 and 1.5 M(cid:12), the two groups are most likely statistically
similar (K-S p-value of 0.06), with the same average value (0.04
in both cases) but a lower dispersion for the field dwarfs. Fi-
nally, comparing the hosts of massive planets with mass above
1.5 M(cid:12) with giant stars of similar mass from Alves et al. (2015),
we also do not find strong hints to refute the hypothesis that the
two groups may come from the same parent distribution (K-S
7 As implemented in the scikit-learn package (Pedregosa et al.
2011).
p-value of 0.03). In this case, the average value of [Fe/H] distri-
bution of the field stars has higher values than the one found for
the host stars of the more massive planets. Again, the standard
deviation has lower values.
We did the same comparison for the hosts of the lower mass
planets (Mpl<4 MJup). In that case, as expected, planet hosts are
always strongly significantly more metal rich than field stars, in
all stellar mass regimes compared. The K-S p-values are very
low, ranging from 3 10−3 for the most massive stars, down to
10−7 and 10−8 for the other stellar mass regimes. The metal-
licity distribution of the low-mass planet hosts thus follows the
usual metallicity-giant planet frequency correlation (Santos et al.
2004; Fischer & Valenti 2005).
3.3. Other planet properties
We further tested whether any other planet property was signif-
icantly different between the two populations of planets (more
and less massive than 4 MJup). For instance, the orbital periods
of the higher mass planet population are indeed longer: 613 ver-
sus 566 days if all stars are considered. These differences are,
however, not statistically significant (K-S p-value of 0.33).
It
is worth mentioning, however, that as already pointed out (e.g.
Adibekyan et al. 2013), lower metallicity stars host, on average,
planets in longer orbital periods. As seen above, the lower-mass
planet population also has higher metallicity, on average, a fact
that could by itself explain this offset.
Finally, we also tested if the eccentricity distributions of the
two samples are different. The results show that no statistically
significant differences exist (K-S p-value of 0.25), even if higher
mass planets tend to have slightly higher average values for the
orbital eccentricity when compared with the lower mass plan-
ets in our selection: 0.29 (STD=0.22) and 0.25 (STD=0.14), re-
spectively. This trend may actually be expected from theoretical
models of planet-disk interaction (Bitsch et al. 2013), as already
discussed in Adibekyan et al. (2013).
4. Discussion
The results presented above suggest that giant planets with
masses above and below ∼4 MJup may represent two different
populations. The data we analyse shows that stars with planets
more massive than 4 MJup orbit stars that are on average more
metal-poor. This trend is statistically significant for the more
massive stars (M > 1.5M(cid:12)), even it if is also observed in all
the analysed stellar mass regimes. We also show that planets
with Mpl>4 MJup orbit stars that span a wider metallicity range
than stars with lower mass planets, and have [Fe/H] distributions
more similar to the average field stars of similar mass.
The fact that the metallicity distribution for the stars with the
more massive planets is lower is intriguing. Metallicity is known
to be intimately related to the frequency of giant planets, a result
coming from both observations and theoretical models based on
the core-accretion paradigm (e.g. Santos et al. 2004; Fischer &
Valenti 2005; Mordasini et al. 2012a). More metal-rich stars may
also be able to form higher mass planets, even if this trend is not
necessarily strong (see Mordasini et al. 2012a, and their Figs. 4
and 5). The results presented here thus seem to contradict these
expectations.
In the context of the core-accretion paradigm, Kennedy &
Kenyon (2008) suggested that the frequency of planets is an in-
creasing function of stellar mass, at least up to ∼3 M(cid:12). This
result is supported by observational evidence (e.g. Reffert et al.
Article number, page 3 of 6
Fig. 2. Metallicity distribution for stars with giant planets in different mass regimes. Upper left: The metallicity distribution is shown for all
stars. Remaining three panels: The metallicity distribution is shown for three different stellar mass regimes.
Table 1. Comparison of the metallicity distributions of the stars with planets in the two mass regimes and values for the solar neighbourhood
populations (for dwarfs/giants).
Sample
All stars
M(cid:63) ≥1.5 M(cid:12)
M(cid:63) ≤1.0 M(cid:12)
1.0 M(cid:12)<M(cid:63)<1.5 M(cid:12)
Mpl > 4 MJup
N <[Fe/H]> STD
0.28
96
0.22
30
48
0.26
0.34
18
−0.04
−0.14
0.04
−0.10
Mpl < 4 MJup
<[Fe/H]> STD K-S p-value
0.07
−0.02
0.13
0.06
5 10−4
3 10−3
0.06
0.06
0.19
0.16
0.20
0.15
N
174
48
89
37
N
241
177
397
Solar neighbourhood
<[Fe/H]>
−0.11/−0.08
−0.08
0.04
−0.16
0.16
0.17
0.24
STD
0.24/0.18
2015, and references within). Higher mass stars are also known
to have higher mass disks (Natta et al. 2000) that are likely ca-
pable of forming higher mass giant planets. It could thus be that
the tendency for the most massive stars in our sample to host
more massive planets may be explained by the simple fact that
their massive disks were able to form planets with higher masses,
even if their metal content was lower. This could explain, in the
context of the core-accretion paradigm, the existence of two pop-
ulations of planets.
If real, however, this model would have to explain the reason
for the proposed change in regime around ∼4 MJup. Further-
more, it would have to explain why in all stellar mass regimes
studied, the metallicity distributions of the hosts of planets with
mass > 4 MJup are always compatible with the field dwarf distri-
bution (even if the [Fe/H] spread is always higher in the planet
hosts), opposite to what is observed if we only consider hosts
of the lower mass planets. For stars with mass below 1 M(cid:12), we
find that planet hosts are more metal rich than field stars (even
if the K-S p-value shows a value that is marginally significant).
Article number, page 4 of 6
However, no difference, or even the inverse trend, is found when
considering stars with mass above this limit.
To further explore this, we compared the index M(cid:63) 10[Fe/H]
of the stars with higher and lower mass planets. Assuming that
the M(cid:63) is correlated with the mass of the disk, this index is ex-
pected to measure the total amount of heavy material in a disk
and thus be more directly related to the planet formation effi-
ciency in the core-accretion scenario. Comparing the two indices
for higher and lower mass planets and using all the stars in our
sample we conclude that planets with mass above 4 MJup have
systematically lower values. A K-S test provides a p-value of
0.08. This result suggests that the increase in stellar mass only
partially compensates for the difference in metallicity. However,
the exact relation between disk mass and stellar mass is not fully
clear (e.g. Andrews et al. 2013), and may have an impact on
these results.
An alternative explanation for the observed populations calls
for the disk instability process (e.g. Boss 1997).
It has been
shown that planets formed by disk instability should in principle
have higher masses and be easier to form around stars with more
N. C. Santos et al.: Observational evidence for two distinct giant planet populations
objects above and below the "brown-dwarf desert". This was
proposed to reflect the existence of two different populations that
were likely formed by different processes: disk instability for the
lower mass brown dwarfs and cloud fragmentation (as "normal"
stars) for the higher mass brown dwarfs (see Ma & Ge 2014).
Although the metallicity-giant planet frequency correlation
is a well-established fact when dealing with dwarf stars, the ex-
istence of such a correlation is still a matter of debate for giant
stars. Indeed, some results suggest that this correlation may even
not be present or could be weaker than that found for the dwarfs
(see discussions in Pasquini et al. 2007; Mortier et al. 2013; Mal-
donado et al. 2013; Reffert et al. 2015). As we have also seen, the
most massive stars in our sample (>1.5 M(cid:12)) are also evolved. In
the scenario mentioned above, the existence of two populations
of giant planets that is observed for the higher mass stars may
be deeply related to the possibility that evolved stars (more mas-
sive on average) also do not show a very clear metallicity-giant
planet correlation.
A word of caution to say that this result is based on the as-
sumption that all planets in the sample are bona fide. This may
not be always the case, especially for planets orbiting giant stars
(see discussion in Reffert et al. 2015).
If these results are confirmed as new planets are detected,
the discussion presented here may give a new important insight
into the giant planet formation process. They also show that
the study of giant planets is still of great importance, even if
the focus of exoplanet research is moving towards the study
of their low-mass counterparts. Giant planets discovered with
the GAIA mission, whose sensitivity will allow the detection of
thousands of giant planets in long period orbits around stars of
different mass(Sozzetti et al. 2001) may shed significant light
into this case. The study of the mass-radius relation of giant
planets as carried out with missions like CHEOPS (Fortier et al.
2014), TESS (Ricker et al. 2010), or PLATO (Rauer et al. 2014),
as well as of their atmospheric composition using new ground-
and space-based instruments such as JWST and HIRES@E-ELT
(e.g. Greene et al. 2016; Marconi et al. 2016), may also bring
new constraints on the processes of planet formation (Fortney
et al. 2008; Mordasini et al. 2012b, 2016).
Acknowledgements. This work was supported by Fundação para a Ciência
e a Tecnologia (FCT, Portugal) through the research grant through national
funds and by FEDER through COMPETE2020 by grants UID/FIS/04434/2013
& POCI-01-0145-FEDER-007672, PTDC/FIS-AST/1526/2014 & POCI-01-
0145-FEDER-016886 and PTDC/FIS-AST/7073/2014 & POCI-01-0145-
FEDER-016880. P.F., S.B., N.C.S. e S.G.S. acknowledge support from FCT
IF/01037/2013CP1191/CT0001,
through Investigador FCT contracts nr.
IF/01312/2014/CP1215/CT0004,
and
IF/00028/2014/CP1215/CT0002. V.A. and E.D.M. acknowledge support from
FCT through Investigador FCT contracts nr. IF/00650/2015/CP1273/CT0001,
IF/00849/2015/CP1273/CT0003;and by the fellowship SFRH/BPD/70574/2010,
SFRH/BPD/76606/2011 funded by FCT and POPH/FSE (EC). PF further ac-
knowledges support from Fundação para a Ciência e a Tecnologia (FCT) in the
form of an exploratory project of reference IF/01037/2013CP1191/CT0001.
ACSF is supported by grant 234989/2014-9 from CNPq (Brazil). J.P.F. acknowl-
edges support from FCT through the grant reference SFRH/BD/93848/2013.
IF/00169/2012/CP0150/CT0002,
References
Adibekyan, V. Z., Figueira, P., Santos, N. C., et al. 2013, A&A, 560, A51
Alves, S., Benamati, L., Santos, N. C., et al. 2015, MNRAS, 448, 2749
Andrews, S. M., Rosenfeld, K. A., Kraus, A. L., & Wilner, D. J. 2013, ApJ, 771,
129
Bashi, D., Helled, R., Zucker, S., & Mordasini, C. 2017, ArXiv e-prints
Beaugé, C. & Nesvorný, D. 2013, ApJ, 763, 12
Bitsch, B., Crida, A., Libert, A.-S., & Lega, E. 2013, A&A, 555, A124
Bonfils, X., Delfosse, X., Udry, S., et al. 2013, A&A, 549, A109
Boss, A. P. 1997, Science, 276, 1836
Article number, page 5 of 6
Fig. 3. Planet mass vs. stellar metallicity plot for our sample stars. In
the bottom panel we show the position of the two populations that result
from our Gaussian mixture analysis (see text for more details).
massive disks (e.g. Rafikov 2005; Nayakshin 2017). Further-
more, it has been suggested that gravitational instability may be
more efficient in more metal-poor disks (or at least not so metal-
licity dependent as the core-accretion process), forming planets
in longer period orbits, and potentially acting as an accelerator
of the core formation (Boss 2002; Cai et al. 2006).
In this scenario, the observational result presented above
could be interpreted as showing the existence of two separate
populations of giant planets formed by different physical pro-
cesses. On the one hand, the lower mass planets (here defined
with a tentative mass below 4 MJup) are formed by the core-
accretion process and are more prevalent around more metal-rich
stars. On the other hand, the more massive planets, whose for-
mation process is mainly done through a gravitational instability
process or a process where disk instability has played a role (Cai
et al. 2006). This second population is less sensitive to the stellar
metallicity. If confirmed, this scenario would imply that above
∼4 MJup the formation of giant planets is no longer dominated
by the core-accretion process. An overlap of the two popula-
tions likely exists, however. A deeper analysis with an increase
in the number of planets in the samples is needed to confirm or
refine this value.
In this context it is relevant to add that studies of stars with
brown-dwarf companions have shown that these have metallic-
ity distributions that are very similar to the solar neighbour-
hood stars (Mata Sánchez et al. 2014; Maldonado & Villaver
2017). The existence of a so-called "brown-dwarf desert", a
pronounced scarcity of companions around Sun-like stars with
mass around ∼30-50 M Jup (e.g. Sahlmann et al. 2011; Ma & Ge
2014), strongly suggests, however, that the higher mass planets
are not likely to be the low-mass end of the "stellar" distribu-
tion. Furthermore, recent results (Maldonado & Villaver 2017)
also suggest that a different metallicity distribution may exist for
-1.0-0.8-0.6-0.4-0.20.00.20.40.6[Fe/H]0246810121416Planetmass[Jupitermasses]-1.0-0.8-0.6-0.4-0.20.00.20.40.6[Fe/H]0246810121416Planetmass[Jupitermasses]223
Boss, A. P. 2002, ApJ, 567, L149
Cai, K., Durisen, R. H., Michael, S., et al. 2006, ApJ, 636, L149
Dawson, R. I. & Murray-Clay, R. A. 2013, ApJ, 767, L24
Fischer, D. A. & Valenti, J. 2005, ApJ, 622, 1102
Fortier, A., Beck, T., Benz, W., et al. 2014, in Society of Photo-Optical Instru-
mentation Engineers (SPIE) Conference Series, Vol. 9143, Society of Photo-
Optical Instrumentation Engineers (SPIE) Conference Series, 2
Fortney, J. J., Marley, M. S., Saumon, D., & Lodders, K. 2008, ApJ, 683, 1104
Gonzalez, G. 1997, MNRAS, 285, 403
Greene, T. P., Line, M. R., Montero, C., et al. 2016, ApJ, 817, 17
Gupta, M. R. & Chen, Y. 2011, Foundations and Trends in Signal Processing, 4,
Ida, S. & Lin, D. N. C. 2004, ApJ, 616, 567
Johnson, J. A., Butler, R. P., Marcy, G. W., et al. 2007, ApJ, 670, 833
Kennedy, G. M. & Kenyon, S. J. 2008, ApJ, 673, 502
Lovis, C. & Mayor, M. 2007, A&A, 472, 657
Ma, B. & Ge, J. 2014, MNRAS, 439, 2781
Maldonado, J. & Villaver, E. 2017, ArXiv e-prints
Maldonado, J., Villaver, E., & Eiroa, C. 2013, A&A, 554, A84
Marconi, A., Di Marcantonio, P., D'Odorico, V., et al. 2016, in Proc. SPIE, Vol.
9908, Society of Photo-Optical Instrumentation Engineers (SPIE) Conference
Series, 990823
Mata Sánchez, D., González Hernández, J. I., Israelian, G., et al. 2014, A&A,
Mayor, M., Lovis, C., & Santos, N. C. 2014, Nature, 513, 328
Mordasini, C., Alibert, Y., Benz, W., Klahr, H., & Henning, T. 2012a, A&A,
Mordasini, C., Alibert, Y., Georgy, C., et al. 2012b, A&A, 547, A112
Mordasini, C., van Boekel, R., Mollière, P., Henning, T., & Benneke, B. 2016,
566, A83
541, A97
ApJ, 832, 41
Mortier, A., Santos, N. C., Sousa, S. G., et al. 2013, A&A, 557, A70
Natta, A., Grinin, V., & Mannings, V. 2000, Protostars and Planets IV, 559
Nayakshin, S. 2017, PASA, 34, e002
Nelson, B. E., Ford, E. B., & Rasio, F. A. 2017, ArXiv e-prints
Ngo, H., Knutson, H. A., Hinkley, S., et al. 2016, ApJ, 827, 8
Pasquini, L., Döllinger, M. P., Weiss, A., et al. 2007, A&A, 473, 979
Pedregosa, F., Varoquaux, G., Gramfort, A., et al. 2011, J. Mach. Learn. Res.,
Rafikov, R. R. 2005, ApJ, 621, L69
Rauer, H., Catala, C., Aerts, C., et al. 2014, Experimental Astronomy, 38, 249
Reffert, S., Bergmann, C., Quirrenbach, A., Trifonov, T., & Künstler, A. 2015,
12, 2825
A&A, 574, A116
Ricker, G. R., Latham, D. W., Vanderspek, R. K., et al. 2010, in Bulletin of
the American Astronomical Society, Vol. 42, American Astronomical Society
Meeting Abstracts #215, #450.06
Sahlmann, J., Ségransan, D., Queloz, D., et al. 2011, A&A, 525, A95
Santos, N. C., Israelian, G., & Mayor, M. 2001, A&A, 373, 1019
Santos, N. C., Israelian, G., & Mayor, M. 2004, A&A, 415, 1153
Santos, N. C., Sousa, S. G., Mortier, A., et al. 2013, A&A, 556, A150
Schneider, J., Dedieu, C., Le Sidaner, P., Savalle, R., & Zolotukhin, I. 2011,
A&A, 532, A79
533, A141+
Sousa, S. G., Santos, N. C., Israelian, G., Mayor, M., & Udry, S. 2011, A&A,
Sozzetti, A., Casertano, S., Lattanzi, M. G., & Spagna, A. 2001, A&A, 373, L21
Torres, G., Andersen, J., & Giménez, A. 2010, A&ARv, 18, 67
Article number, page 6 of 6
|
1512.03535 | 1 | 1512 | 2015-12-11T06:23:34 | The Inner Debris Structure in the Fomalhaut Planetary System | [
"astro-ph.EP"
] | Fomalhaut plays an important role in the study of debris disks and small bodies in other planetary systems. The proximity and luminosity of the star make key features of its debris, like the water ice-line, accessible. Here we present ALMA cycle 1, 870 \mu m (345 GHz) observations targeted at the inner part of the Fomalhaut system with a synthesized beam of 0.45"x0.37" (~3 AU linear resolution at the distance of Fomalhaut) and a rms of 26 \mu Jy/beam. The high angular resolution and sensitivity of the ALMA data enable us to place strong constraints on the nature of the warm excess revealed by Spitzer and Herschel observations. We detect a point source at the star position with a total flux consistent with thermal emission from the stellar photosphere. No structures that are brighter than 3\sigma\ are detected in the central 15 AU x 15 AU region. Modeling the spectral energy distribution using parameters expected for a dust-producing planetesimal belt indicates a radial location in the range ~8-15 AU. This is consistent with the location where ice sublimates in Fomalhaut, i.e., an asteroid-belt analog. The 3\sigma\ upper limit for such a belt is <1.3 mJy at 870 \mu m. We also interpret the 2 and 8-13 \mu m interferometric measurements to reveal the structure in the inner 10 AU region as dust naturally connected to this proposed asteroid belt by Poynting-Robertson drag, dust sublimation, and magnetically trapped nano grains. | astro-ph.EP | astro-ph | accepted for publication in ApJ
Preprint typeset using LATEX style emulateapj v. 5/2/11
THE INNER DEBRIS STRUCTURE IN THE FOMALHAUT∗ PLANETARY SYSTEM
Kate Y. L. Su1,6, George H. Rieke1, Denis Defr`ere1, Kuo-Song Wang2, Shih-Ping Lai3,6, David J. Wilner4, Rik
van Lieshout5, and Chin-Fei Lee2
1Steward Observatory, University of Arizona, Tucson, AZ 85721, USA; [email protected]
2Institute of Astronomy and Astrophysics, Academia Sinica, P.O. Box 23-141, Taipei 106, Taiwan
3Institute of Astronomy, National Tsing Hua University (NTHU), Hsinchu 30013, Taiwan
5
1
0
2
c
e
D
1
1
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
5
3
5
3
0
.
2
1
5
1
:
v
i
X
r
a
5Anton Pannekoek Institute for Astronomy, University of Amsterdam, Science Park 904, 1098 XH Amsterdam, The Netherlands
4Harvard-Smithsonian Center for Astrophysics 60 Garden Street, Cambridge, MA 02138, USA and
ABSTRACT
Fomalhaut plays an important role in the study of debris disks and small bodies in other planetary
systems. The proximity and luminosity of the star make key features of its debris, like the water
ice-line, accessible. Here we present ALMA cycle 1, 870 µm (345 GHz) observations targeted at the
inner part of the Fomalhaut system with a synthesized beam of 0.(cid:48)(cid:48)45×0.(cid:48)(cid:48)37 (∼3 AU linear resolution at
the distance of Fomalhaut) and a rms of 26 µJy beam−1. The high angular resolution and sensitivity
of the ALMA data enable us to place strong constraints on the nature of the warm excess revealed
by Spitzer and Herschel observations. We detect a point source at the star position with a total flux
consistent with thermal emission from the stellar photosphere. No structures that are brighter than
3σ are detected in the central 15 AU×15 AU region. Modeling the spectral energy distribution using
parameters expected for a dust-producing planetesimal belt indicates a radial location in the range
∼8 -- 15 AU. This is consistent with the location where ice sublimates in Fomalhaut, i.e., an asteroid-
belt analog. The 3σ upper limit for such a belt is <1.3 mJy at 870 µm. We also interpret the 2
and 8 -- 13 µm interferometric measurements to reveal the structure in the inner 10 AU region as dust
naturally connected to this proposed asteroid belt by Poynting-Robertson drag, dust sublimation, and
magnetically trapped nano grains.
Subject headings: circumstellar matter -- radio: stars, planetary systems -- stars: individual (Fomal-
haut)
1. INTRODUCTION
The Fomalhaut debris disk has become the prototype
for understanding the complexity of disk structures and
their relationship to planetary systems. Due to its prox-
imity (7.7 pc) and luminosity (17.4 L(cid:12)), many key fea-
tures of its debris are easily accessible through optical,
infrared and millimeter ground- and space-based facili-
ties. The prominent planetesimal ring located at ∼140
AU, analogous to the Kuiper belt in our Solar System,
is particularly well-studied (Holland et al. 1998; Kalas et
al. 2005; Acke et al. 2012; Ricci et al. 2012). The sharp
boundaries of this cold (∼50 K) ring suggest gravita-
tional shaping by unseen planets (Quillen 2006; Chiang
et al. 2009; Boley et al. 2012). Intense collisions among
the planetesimals in this ring grind large (parent) bodies
into fine dust, forming a disk halo composed of ∼µm size
grains outside the cold ring (Acke et al. 2012; Kalas et
al. 2013). The Fomalhaut disk has a warm (∼150 K),
unresolved component near the star, first indicated by
Spitzer 24 µm imaging (Stapelfeldt et al. 2004) and later
confirmed by Herschel imaging (Acke et al. 2012). This
warm excess has been suggested as an ice-line planetesi-
mal belt, analogous to the asteroid belt in our Solar Sys-
tem (Su et al. 2013). Furthermore, the Fomalhaut sys-
tem is thought to have a very hot (∼1500 K) exozodiacal
∗
Fomalhaut is a triple system; here we refer the Fomalhaut
planetary system as the one around the primary star Fomalhaut
A.
6 Visiting Scholar, Institute of Astronomy and Astrophysics,
Academia Sinica, P.O. Box 23-141, Taipei 106, Taiwan
dust population at ∼0.1 AU, revealed by ground-based
interferometry at 2 µm (K-band) (Absil et al. 2009). In-
terferometric data at ∼10 µm detect a hot component
near ∼1 -- 2 AU (Mennesson et al. 2013; Lebreton et al.
2013). Therefore, the Fomalhaut system is the only sys-
tem known to possess all five debris zones (very hot, hot,
warm, cold, and halo dust) defined by Su & Rieke (2014).
In addition to the debris system around Fomalhaut, a
combination of imaging where the warm component can
be spatially separated from the cold component and mid-
infrared spectroscopy has revealed similar inner warm
components in other nearby systems, for example, Eri:
(Backman et al. 2009; Greaves et al. 2014), HR 8799: (Su
et al. 2009; Matthews et al. 2014), Vega: (Su et al. 2013).
∼20% of unresolved debris disks show such warm compo-
nents in their disk spectral energy distributions (SEDs)
(Morales et al. 2011; Ballering et al. 2013; Chen et al.
2014; Kennedy & Wyatt 2014). Leftover parent bodies
near the star generally have higher collisional velocities,
i.e., shorter dynamical time scales, compared to those of
more distant ones, as suggested in the fading of the warm
components after a few hundred Myr (e.g., G´asp´ar et al.
2013). These components may be analogous to our as-
teroid belt but with far greater mass; however, compared
to the cold Kuiper-belt analogs they are much harder to
characterize in exoplanetary systems. Identification and
characterization of such warm components require very
accurate extrapolation of the stellar spectrum and pre-
cise subtraction.
Not surprisingly, although it can be considered the pro-
totype, the nature of the Fomalhaut warm component is
2
Su et al.
controversial. Acke et al. (2012) propose that it origi-
nates from free-free emission of a postulated stellar wind
that also creates the hot excess revealed in near-infrared
interferometry. However, Su et al. (2013) draw from all
the resolved images and the spectroscopy to argue that
it is a planetesimal ring placed near the ice line and pre-
sumably created by processes occurring in other debris
systems as well. In that case the Fomalhaut debris disk
resembles the debris structures in our Solar System --
a dense Kuiper belt whose inner edge is maintained by
massive planets (Liou & Zook 1999) and a more ten-
uous asteroid belt containing dust structure (zodiacal
cloud) determined by both the giant and terrestrial plan-
ets (Dermott et al. 1994).
There is a tentative detection of an excess in the ALMA
cycle 0 image of Fomalhaut (Boley et al. 2012), but less
than the prediction by the free-free model, and the star
was at the edge of the primary beam so the excess flux
is subject to the large uncertainty of the primary beam
correction. Here we present a high angular resolution
and sensitivity ALMA cycle 1 image centered at the star
position, and provide a strong constraint on the amount
of excess emission in the inner part of the Fomalhaut
system.
Section 2 of this paper describes the details of the
ALMA observation and data reduction.
In Section 3,
we discuss the 870 µm measurement near the star, evalu-
ate the expected photospheric value of Fomalhaut at this
wavelength, and rule out any compact free-free emission
originated from the star as the possible source of excess
emission detected by Spitzer and Herschel.
In Section
4.1 we review the properties of the inner 20 AU region
in Fomalhaut by assessing all available excess measure-
ments and discussing their uncertainties. We conclude
that a warm excess at the ice sublimation temperature
(∼150 K, i.e., an asteroid-belt analog) is robustly de-
tected, independent of the uncertainty of photospheric
subtraction and the presence of a tentative, hot (∼500
K) excess.
In Section 4.2, we present SED models for
the ice-line asteroid belt around Fomalhaut and set con-
straints on its total flux at 870 µm, and we also show
that a 13-AU narrow belt might be detected at (cid:46)2σ lev-
els in the cycle 1 ALMA data. In Section 4.3, we propose
a new interpretation for the hot excess detected by the
Keck Interferometric Nuller (KIN) via linking the pro-
posed asteroid belt to the dust structure in the inner 10
AU region. Additional validation in terms of the derived
dust mass and the sensitivity of the adapted grain param-
eters to our model are discussed in Section 5, followed by
our conclusions in Section 6.
2. OBSERVATIONS AND DATA REDUCTION
Fomalhaut was observed with the ALMA 12-m array in
Cycle 1 (program 2012.1.00238.S, PI: Lai) in five execu-
tion blocks at band 7. Table 1 summarizes the detailed
information for each execution including the bandpass,
amplitude, phase calibrators, and the estimated precip-
itable water vapour (PWV). The data taken on the sec-
ond execution of 2014 May 21 did not pass the quality as-
surance procedure, thus are not included in the imaging
and analysis study. A total on-source time of 172 minutes
was achieved. The phase center is located at the star po-
sition (RA=22:57:39.409, Dec=−29:37:22.423) corrected
for proper motion at the epoch of the observation. Four
2-GHz wide spectral windows centered at 337.987 GHz,
339.924 GHz, 350.002 GHz and 352.002 GHz were used
to measure the continuum emission with dual linear po-
larizations from the target source. The primary beam
size (the half-power width of the Gaussian beam) is 18(cid:48)(cid:48)
at 345 GHz. The overall uv-coverage ranges from 23.0 kλ
to 737.7 kλ. Thus, our data are not sensitive to smooth
emission structures larger than ∼9(cid:48)(cid:48). The width of the
outer belt is estimated to be ∼13 -- 19 AU (Boley et al.
2012), which is mostly resolved out by our high resolu-
tion (∼3 AU in linear scale); therefore, our data only
show a very faint trace of the outer belt along the minor-
axis direction. The quasar J2258−279 was used as the
amplitude calibrator in all execution blocks. The flux
uncertainty is estimated to be 10%.
All the data reduction and imaging processes were car-
ried out with CASA 4.2.1 (McMullin et al. 2007). The
final continuum image was obtained by performing multi-
frequency-synthesis (mfs) imaging. Clean boxes of the
expected cold belt location (generated with reference to
the HST image) and the central 5(cid:48)(cid:48) for the warm belt,
were used in the image deconvolution process. A final
image with effective bandwidth of 8 GHz centered at 345
GHz was generated. The synthesized beam is 0.(cid:48)(cid:48)45×0.(cid:48)(cid:48)37
with position angle (P.A.) of 82.3◦. The rms noise in the
final mfs image is 26 µJy beam−1.
3. ALMA OBSERVATIONAL RESULTS
3.1. 870 µm Measurement
Figure 1(a) shows the central 6(cid:48)(cid:48)×6(cid:48)(cid:48) region of the map.
A bright source, detected at the star position after taking
the star's proper motion into account, has a full-width-
half-maximum (FWHM) of 0.(cid:48)(cid:48)46×0.(cid:48)(cid:48)38 at P.A. of 86.4◦,
consistent with being the star (also see Sec 3.2). To un-
cover any faint structure that might be present around
the star, we fitted a single point source model to the data
and subtracted it from the original data in the uv plane,
then generated the residual map. We used uvmodelfit in
CASA to determine the best-fit parameters for the point
source: a total flux of 1.789±0.037 mJy. The quoted
uncertainty is based on the fit, i.e., this number does
not include other errors from flux calibration. The resid-
ual map is shown Figure 1(b). No structures that are
brighter than 3σ are detected in the 2(cid:48)(cid:48)×2(cid:48)(cid:48) (15 AU×15
AU) region around the star after subtracting the best-fit
point source model. Assuming the inner belt has a sim-
ilar orientation as the outer belt, one would expect the
disk ansae to be along the P.A. of 156◦. There are two
∼2σ blobs along that angle: one is ∼0.(cid:48)(cid:48)4 (3 AU) north-
west of the star, and the other is ∼1.(cid:48)(cid:48)8 (14 AU) south-east
of the star. Based on our SED model (see Section 4.2),
the likely location of the inner belt is at radii of ∼8 -- 15
AU, suggesting the south-east blob could be part of the
inner belt. Assuming the inner belt is also inclined by
66◦ like the outer one, but centered at the star (no offset),
the ellipse on Figure 1(b) marks the boundary of a puta-
tive 13-AU inner belt. Interestingly, there are a few 1 -- 2σ
blobs along the expected disk circumference, suggesting
this 13-AU inner belt might be detected at 1-2σ levels
(more discussion is given in Section 4.2). Although the
1-2σ positive blobs along the non-offset, expected 13-AU
disk circumference are intriguing, we note that there are
also many −2σ blobs in the data as well. In summary,
3
ALMA Cycle 1 Observations of Fomalhaut
Table 1
EB
code
X4a7
X11ad
X135f
Xe46
X119f
UT Date
2014 Apr 29
2014 May 21
2014 May 21
2014 May 26
2014 May 26
Passband
J2258−2758
J2258−2758
J2258−2758
J2056−4714
J2258−2758
Calibrators
Flux
J2258−279
J2258−279
J2258−279
J2258−279
J2258−279
Phase
J2250−2806
J2250−2806
J2250−2806
J2258−2758
J2250−2806
# of
anntena
PWV
[mm]
35
32
32
25
25
1.43
1.10
1.24
0.67
0.72
we do not find any significant extended structure around
the star, and the reality of the putative 13-AU belt needs
future confirmation.
3.2. Expected Photosphere Brightness at 870 µm and
A Constraint on the Unresolved Excess
Su et al. (2013) evaluated all available optical to near-
infrared photometry and determined the stellar photo-
spheric output to be 2.96 Jy in the MIPS 24 µm band.
A simple Rayleigh-Jeans extrapolation then yields an es-
timate of 2.2 mJy at 870 µm, significantly brighter (by
∼23%) than the point source measured at the star posi-
tion. The discrepancy likely comes from two sources: (1)
the uncertainty of the ALMA absolute calibration, and
(2) the uncertainty of Rayleigh-Jeans extrapolation.
The accuracy in absolute calibration for ALMA Cycle
2 is set at 10% for Band 7 (J. Mangum, priv. communica-
tion). To verify the calibration, we used the solar analog
method (Johnson 1965; Rieke et al. 2008) by comparing
measurements of α Cen A with those of the Sun. We
did this in two parts. First, we verified α Cen A is a so-
lar analog by comparing its colors with those of the Sun.
Then, we computed the submillimeter flux of α Cen A us-
ing the solar relation between measured 24 µm and sub-
millimeter. We derived colors of V − Ks=1.531±0.023,
V − [24]=1.584±0.027, and Ks − [24]=0.053±0.032 us-
ing the available K-band measurement from Engels et al.
(1981), V -band from Alekseeva et al. (1996), Mermilliod
(1991) and Hipparcos, and 24 µm measurement of 30.84
Jy from Wiegert et al. (2014) after proper photometric
system transformation and calibration to the zero points
defined by Rieke et al. (2008). These colors are virtually
identical to those of the Sun and establish α Cen A as a
valid solar analog star for the purpose of calibration.
We then used measurements of the solar brightness
temperature to connect the mid-infrared measurements
to the ALMA calibration at 870 µm. We adopt the
brightness temperature at 24 µm to be 4625 K (Model
M in Vernazza et al. 1976), and the brightness tempera-
ture at 870 µm to be 5470 K (Loukitcheva & Nagnibeda
2000). Taking 30.84±0.76 Jy as the 24 µm flux of α
Cen A (Wiegert et al. 2014), the scaling from the solar
brightness temperatures suggests α Cen A has a flux den-
sity of 28.9 mJy at 872 µm. The error associated with
this number is not straightforward to determine, but can
be estimated as the quadratically combined error of 7%
(±300 K (6%) in the solar 870 µm brightness temper-
ature, ∼2% in the solar 24 µm brightness temperature,
and 2.5% of the α Cen A 24 µm measurement). The mea-
surement of 26.1±0.2 mJy for α Cen A from Liseau et
al. (2015) then suggests that the solar analog calibration
is brighter than the ALMA measurement by 10.7±7%,
which confirms the 10% ALMA calibration uncertainty
estimate is plausible.
If we adjust the measurement of Fomalhaut to this al-
ternative calibration, we find a flux density of 1.98 mJy
(instead of 1.789 mJy), still somewhat fainter (by ∼11%)
than the Rayleigh-Jeans extrapolation from 24 µm. We
conclude that radiative transfer effects may cause a drop
in the brightness temperature of the star, similar to the
behavior of the Sun, which has a brightness temperature
≥5800 K across the visible (Vernazza et al. 1976) but
of only 5470 K at 870 µm. The result that Fomalhaut
is slightly fainter than the Rayleigh-Jeans extrapolation
value is robust, and this independent calibration of our
measurement also rules out any compact free-free emis-
sion originating from the star as the possible source of
excess emission detected by Spitzer and Herschel.
4. ANALYSIS
4.1. Properties of Inner 20 AU Excess in Fomalhaut
In this subsection, we review all the available measure-
ments that can be used to constrain the excess properties
of the inner 20 AU region in Fomalhaut. These measure-
ments can be categorized in two classes: (I) resolved mea-
surements where emission from the cold belt is mostly
excluded, and (II) unresolved measurements at the wave-
lengths where the cold belt contributes much less emis-
sion. In Category (I), the ground-based interferometric
measurements suggest a K-band excess of 0.88%±0.12%
(Absil et al. 2009, hereafter the VLTI measurement) and
8 -- 13 µm excess of 0.35%±0.10% (Mennesson et al. 2013,
hereafter the KIN measurements) relative to the stel-
lar photosphere. These interferometric observations can
very effectively isolate the region that is very close to the
star ((cid:46)2 AU in K-band and (cid:46)6 AU at 8 -- 13 µm) from
the contribution of the stellar photosphere. In addition,
the excess emission at 24 and 70 µm inferred from the
resolved images of Fomalhaut by Spitzer and Herschel
also belongs to the first category. Since the star and any
unresolved excess are both contained within the instru-
ment's point spread function (PSF), the excess estimate
does depend on the expected photospheric values.
In determining the unresolved excess, an accurate
value for the photospheric flux is essential; it can be de-
termined at ∼2% levels by fitting the stellar atmospheric
models to the optical to near infrared photometry (Su
et al. 2005; Engelbracht et al. 2007). At 24 µm, an ad-
ditional unresolved source of 0.6±0.2 Jy was found in
the early reduction of the MIPS data (Stapelfeldt et al.
2004). We reduced the original data with the final in-
house reduction pipeline and calibration, and employed
the PSF subtraction method. We find the unresolved
4
Su et al.
Figure 1. (a) Cycle 1 ALMA map of the central 6(cid:48)(cid:48)×6(cid:48)(cid:48) region centered at the stellar position in the Fomalhaut system. The synthesized
beam, shown as a white ellipse, has a FWHM of 0.(cid:48)(cid:48)45×0.(cid:48)(cid:48)37, ∼3 AU linear resolution at the distance of Fomalhaut (7.7 pc). Our cycle 1
observations reach a rms of 26 µJy beam−1. (b) Star-subtracted residual map. No structures that are brighter than 3σ are detected in
the 2(cid:48)(cid:48)×2(cid:48)(cid:48) (15 AU×15 AU) central region. The expected disk circumference of a 13-AU narrow belt, which has the same inclination and
position angles as the outer cold belt but centered at the star, is marked as the dashed ellipse. A few ∼2σ blobs are found along the ellipse.
excess at 24 µm is 0.64±0.13 Jy, consistent with the old
PSF fitting photometry (±1.5% level) by Marengo et al.
(2009) for Fomalhaut and Sirius in the IRAC bands. The
value (but with a smaller uncertainty). Acke et al. (2012)
color difference between Fomalhaut and Sirius is 2.36
constrained the amount of the unresolved excess by fit-
ting the resolved images with a SED model using complex
mag at IRAC 1, 2, 3 bands, but 2.33 mag at IRAC band
4, suggesting a 3%±1.5% excess at IRAC 8 µm (a very
dust grains (a mixture of compositions proposed by Min
et al. (2011)), and derived a total of 0.17±0.02 Jy for the
broad bandwidth of 36%). Since the IRAC measurement
was derived by PSF fitting, the excess most likely origi-
unresolved excess at 70 µm. Su et al. (2013) employed
the PSF subtraction method, and derived a consistent
nates from the region within 8 AU (in radius) (resolution
of ∼2(cid:48)(cid:48) at IRAC 8 µm).
number, 0.17 Jy with upper- and lower-bound fluxes of
0.58 Jy and 0.136 Jy for the unresolved excess. The reso-
lution at both Spitzer 24 µm and Herschel 70 µm is very
similar (∼6(cid:48)(cid:48)), which places the excess emission mostly
from the region within ∼20 AU in projected stellocentric
distance.
In Category (II), the Spitzer IRS spectrum of the Fo-
malhaut system presented in Su et al. (2013) provides
good measurements in the 10 -- 30 µm region. The IRS
spectrum was extracted and calibrated as a point source
(resolution of ∼6(cid:48)(cid:48)). With a large slit size of 11.(cid:48)(cid:48)1 in
the IRS long-high channel, the flux longward of ∼30
µm might be contaminated by the cold ring. Further-
more, the IRS excess emission is subject to the uncer-
tainty of photospheric subtraction. To account for this,
the uncertainty for the IRS excess emission includes 2%
of the photospheric emission (added quadratically) as
shown in Figure 2. To illustrate the effect of photo-
spheric subtraction, we also present excess spectra by ar-
tificially scaling our best-determined photospheric model
by ±2%. The resultant spectra (shown in Figure 2) have
higher/lower flux in the 10 -- 15 µm region, but both are
within the estimated uncertainty (gray area). We aug-
ment the IRS excess determination with additional ac-
curate1 broadband photometry. We adopt the accurate
1 To obtain infrared excesses at the ∼2% level, using a color
One good way to characterize where the bulk of ex-
cess emission comes from is to determine the dominant
dust temperature in the excess spectrum. This domi-
nant dust temperature simply means that the majority
of the dust grains have a similar dust temperature, which
is not subject to a specific grain model even though this
temperature is usually derived by the blackbody formula.
Fitting the IRS excess emission, the inner warm emission
is best described by a blackbody emission of ∼170 K (see
Figure 2); and the temperature remains the same for the
excess spectrum that was over-subtracted by 2% more of
the photosphere. If the photosphere level is lower by 2%
compared to the nominal level (i.e., under-subtracted),
the resultant IRS excess emission needs at least two tem-
peratures to fit, 167 K and 497 K. Due to the uncer-
tainty of the photospheric subtraction, we cannot rule
out the presence of ∼500 K dust in the system, espe-
cially, the two-temperature fit is a better fit to the IRAC
8 µm point. An alternative fit in this case is a single
blackbody plus a weak (broad) silicate emission feature
difference referenced to a clean, no infrared excess star (like Sir-
ius) can avoid the uncertainty in photospheric subtraction as long
as both the star of interest and the reference star (Sirius) have
a similar spectral type and accurate photometry. Unfortunately,
ALLWISE and AKARI IRC catalogs do not match the latter cri-
terion sufficiently well.
(a)(b)5
Figure 3. SED of the infrared excess for the inner 20 AU region
around Fomalhaut with various models. The data points are the
same as in Figure 2, except for the ALMA Band 7 upper limits.
The measurement from the cycle 0 Band 7 (Boley et al. 2012) is
shown as a filled square with a downward arrow, while the one
from cycle 1 is shown as an orange downward arrow. The model
from Lebreton et al. (2013) is shown as the thick dot-dash line
composed of two belts (thin dotted-dash lines) at ∼0.1 and ∼2 AU.
Our narrow ring (NR) models are aimed to fit the excess emission
near the ice line (asteroid-belt analog) and are constrained by data
points longward of 10 µm.
dius) is ∼2.5 -- 5 µm around a 2.25 M(cid:12) star with a lumi-
nosity of 17.4 L(cid:12). The maximum grain size is set at 1
mm since grains larger than this size contribute negligible
emission in the infrared. We further assume the warm
dust belt is confined in a narrow ring (NR) with a fixed
width of 2 AU. The exact function of the density distribu-
tion in the warm belt has little impact in the output SED
for such a narrow ring; therefore, we adopt a Gaussian
profile for the ring's density distribution. Because we
lack good constraints on the Rayleigh-Jeans side of the
excess emission, the SED models are not well constrained
even with all the assumptions mentioned before. A small
range of ∼8 -- 15 AU with various combinations of mini-
mum grain size and size distribution slope can all provide
satisfactory fits to the observed SED. We illustrate this
degeneracy by presenting two narrow-ring models peaked
at 9 and 13 AU shown in Figure 3, where both models
have the same grain size distribution (5 µm to 1 mm
silicate grains with a size slope of −3.65).
Assuming the inner belt is inclined by 66◦ from face-on
like the outer cold belt, a narrow ring of 9 (13) AU would
cover ∼13 (17) beams in its circumference in our cycle 1
map. Given a rms of 26 µJy beam−1, the corresponding
3σ upper limit for the whole inner belt is 1 mJy and 1.3
mJy at 870 µm for such rings. Our narrow-ring SED
models give a much fainter total flux (see Figure 3), con-
sistent with no detection. As presented in Section 3.1,
there are a few 1 -- 2σ blobs along the expected 13-AU-ring
circumference. We simulate the 870 µm model image of
the 13-AU narrow belt under the same observing depth
and array configuration as the cycle 1 data. Figure 4(a)
shows the synthesized image without including random
thermal noise. We then added the random thermal noise
by adjusting the PWV parameters in simobserve until
the measured noise level reaches 26 µJy beam−1 as in our
cycle 1 observation. We then applied the same procedure
that generated Figure 1(b) to generate a star-subtracted
residual map of the synthesized image shown in Figure
4(b) for comparison. Note that the noise in our simula-
Figure 2. SED of the infrared excess for the inner 20 AU region
around Fomalhaut. The thick green line is from the IRS spec-
trum presented in Su et al. (2013) after the nominal photospheric
subtraction. The gray area represents the uncertainty of the IRS
excess, including 2% of the photospheric subtraction. The thin
orange lines represent the excess IRS spectra if the photosphere
is ±2% higher/lower than the nominal value. The filled squares
represent the broadband excess emission estimated from the re-
solved images, the open square is from the IRAC measurement
(see text for details), and the 3σ upper limit at 160 µm is shown as
a downward arrow. With the nominal and 2% higher photospheric
subtraction, the IRS excess spectra (green and lower orange lines)
are best described by a blackbody emission of ∼170 K for wave-
If the photosphere is lower by 2%,
lengths shorter than 30 µm.
the resultant IRS excess spectrum is best described by a combina-
tion of two blackbody temperatures at 167 K and 497 K (two gray
dotted dash lines).
(see discussion in Section 4.3.2 and Figure 6 for some ex-
amples). Even with this tentative excess emission either
at ∼500 K or a broad feature, the 10 -- 30 µm excess can
only be fitted with a cooler component at 167 K, suggest-
ing the robustness of the warm excess at the temperature
of water ice sublimation (∼150 K), i.e., an asteroid-belt
analog.
4.2. Parameters for the Asteroid Belt Model
We use a simple optically thin SED model to esti-
mate the physical location of the asteroid belt. Our
SED model is based on the assumption that the warm
(∼170 K) excess originates from a planetesimal belt lo-
cated near the water ice line. Therefore, we only use data
longward of ∼10 µm to constrain our fits and do not in-
clude the possible hot ∼500 K component (see Section
4.3 for further discussion).
In this scenario, dust de-
bris is generated through collisions of large parent bod-
ies and cascades down to fine grains (close to the ra-
diation blowout size) with a typical power-law slope of
−3.5∼−3.7 (Wyatt et al. 2011; G´asp´ar et al. 2012). The
radiation blowout size2 depends on the stellar fundamen-
tal parameters (mass and luminosity) and grain proper-
ties (size and density). Using astronomical silicates (Laor
& Draine 1993) with various densities ranging from 1.8 --
3.3 g cm−3 to account for porosity, the blowout size (ra-
3QprL∗
16πc GM∗aρg
2 The grain size can be parametrized by the β value defined as
the ratio of the radiation force and the gravitational force on a
particle, i.e., β =
where Qpr is the radiation pres-
sure efficiency averaged over the stellar spectrum, L∗ and M∗
are the stellar luminosity and mass, G is the gravitational con-
stant, c is light speed, a is the grain radius, and ρg is the grain
density. The blowout size (abl) is defined when β = 0.5,
i.e.,
µm = 1.153 Qpr
abl
ρg
g cm−3
L∗
L(cid:12)
M(cid:12)
M∗
1
10100wavelength (µm)101102103flux density (mJy)One BB of 171 KTwo BBs of 167 K + 497 Kflux contaminatedfrom the cold ringIRS excess after nomial phot. subIRS excess with +/− 2% phot. unc.101001000wavelength (µm)10−1100101102103flux density (mJy)Lebreton+ 2013 modelflux contaminatedfrom the cold ring 9 AU NR model 13 AU NR model6
Su et al.
Figure 4. (a) Synthesized image of our model for a 13-AU narrow belt at the same depth as the cycle 1 observation (details see Section 4.2)
where the synthesized beam is shown as the white ellipse at the left-hand corner. No random thermal noise is included in this simulation.
The contours mark the inner belt at the levels of 20 and 36 µJy beam−1. (b) Synthesized inner belt image with star subtraction. Random
thermal noise is added to the simulated image (a) first, then the same procedure that produced the Figure 1(b) (a point source fitting and
subtraction in the uv plane) was applied to generate this model residual image. The south-east disk ansa appears to be at ∼2σ levels based
on this simulated model.
tion is added randomly; therefore, it is not expected to
coincide with the observation when compared to Figure
1b. Futhermore, we did not include any background con-
fusion as estimated by Oteo et al. (2015) using ALMA
calibration data, i.e., Figure 1b is expected to have some
contribution of faint ((cid:38)3) background galaxies. Our sim-
ulation suggests that the south-east disk ansa might be
detected at ∼2σ levels. Nevertheless, a deeper map is
needed to confirm this putative 13-AU belt.
4.3. Complexity of Inner Debris Distribution -- A New
Interpretation of Interferometric Data
4.3.1. A New Model for the Inner 20 AU Excess in
Fomalhaut
Our proposed asteroid belt (a narrow belt near the wa-
ter ice line, presented in Section 4.2) would not produce
any detectable signal in the KIN measurements because
the dust in this narrow belt is too cold (i.e., too faint) at
8 -- 13 µm, and it is located mostly outside the field of view
of KIN. To reconcile with the KIN measurements, here
we explore a different scenario including a P-R drag-in
component from the asteroid belt and a hot (∼1500 K)
ring produced by the magnetically trapped nano grains
proposed by Rieke et al. (2015, accepted). To roughly
estimate the feasibility of a P-R component from the
proposed asteroid belt, we first compare the typical P-R
timescale with the collisional one in our proposed ice-
line belt3. The collisional time scale is ∼2×104 yr for
a belt at 10 AU with an optical depth of 1×10−4 (see
√
r3GM∗
2τeff
and
3 The collisional time scale is formulated as tcoll =
the P-R time scale is as tPR = cr2
distance, τeff is the optical depth.
4GM∗β where r is the stellocentric
below) around a 2.25 M(cid:12) star. This collisional timescale
is slightly shorter than the P-R timescale, ∼3.6×104 yr
for β =0.5 grains, suggesting some amount of material
can drift inward under the influence of P-R drag, forming
an interior extended disk. Additional material might be
deposited in this region by disintegrating comets; hence
the P-R drag component represents a rough lower limit
but one that can be analyzed without introducing un-
controlled free parameters.
We construct individual SED models for each of the
components according to its spatial constraint and ex-
pected grain population in the drag-in disk. By adjusting
the individual contribution of each component, we then
simultaneously obtain good fits in the KIN measurements
and the overall SED.
The amount of material that can be brought inward
from a dust-producing planetesimal belt due to P-R drag
has been studied analytically (Wyatt 2005) and numer-
ically (van Lieshout et al. 2014). Basically, it depends
strongly on the amount of material (i.e., the collision
rate) in the planetesimal belt. Assuming a single grain
size in the belt and that the collisions are destructive,
the effective optical depth (τeff (r), the vertical optical
depth for a face-on disk) is parametrized as equations
(4) and (5) in Wyatt (2005). To estimate τeff (r0) in
Wyatt's formulae4, we use the observed fractional lu-
minosity (fd) and the relationship, τeff (r0) = 2fdr0
∆r ,
given by Kennedy & Piette (2015) (where r0 and ∆r
4 The analytical model by Wyatt (2005) has very simplified as-
sumptions: grains with one single size and emitting like blackbod-
ies. It is then not straight forward to use the optical depth from a
SED model (a size distribution of grains with imperfect absorption
coefficient) in these analytical formulae.
(a)(b)are the belt's location and width.). The fd values are
∼1.2×10−5 in our SED models presented in Section 4.2;
therefore, τeff (r0) is in the range of ∼1×10−4. For a
planetesimal belt at ∼10 AU around a 2.25 M(cid:12) star
with an initial optical depth of 1×10−4, the value of η0
(equation (5) in Wyatt 2005) is ∼2, and equation (4),
τeff (r) = 1× 10−4[1 + 8(1−(cid:112)r/10AU ]−1, gives the max-
imum amount of material that can spiral inward. van
Lieshout et al. (2014) performed detailed numerical sim-
ulations by including a size distribution of particles in
a collision-dominated planetesimal belt, and found that
the amount of the material due to P-R drag is roughly a
factor of 7 lower than the simple analytical calculation.
As suggested by Kennedy & Piette (2015), one can sim-
ply scale the η0 value by a multiplicative factor k where
k = 1/7 to match the numerical result from van Lieshout
et al. (2014). In addition, van Lieshout et al. (2014) also
found that the size slope in the drag-in component is ex-
pected to be steep with a wavy distribution. Following
the wording suggested in Kennedy & Piette (2015), we
refer to Wyatt's analytical study as the low collision case
and van Lieshout's numerical result as the high collision
case (or collision-dominated). Furthermore, Kobayashi
et al. (2009) suggest a density enhancement (a pile-up
effect) can occur near the dust sublimation radius of a
P-R drag-in disk (i.e., the inner edge of the P-R disk).
As a dust grain drifts close to the sublimation radius and
starts to sublimate (i.e., reducing its size), the radiation
force on the dust grain becomes stronger and temporar-
ily halts its inward migration; therefore, a ring near the
sublimation radius can form. The sublimation radius and
the enhancement factors depend on the grain composi-
tion. Refractory materials like silicate and carbon grains
can give an enhancement factor up to ∼4 -- 6 (Kobayashi
et al. 2011).
Based on the theoretical models described above, we
construct the SED of the P-R component in two parts.
The first part is the drag-in disk component that has
a constant surface density, starting from ∼10 AU (the
inner boundary of the asteroid belt) to ∼0.23 AU (the
sublimation radius for silicate-like grains when they reach
∼1300 K), and is composed of a population of astronom-
ical silicates in a power-law size distribution with a slope
of −5.5, and with sizes ranging from 3 to 10 µm. The
choices of grain size parameters follow the recipe devel-
oped by Wyatt (2005) and van Lieshout et al. (2014)
where grains with β = 0.5 are the most dominant sizes
as the product of collisional cascades, hence, they con-
tribute the most emission from the drag-in disk (see Fig-
ure 5 in van Lieshout et al. 2014). The second part is
for the density enhancement near the silicate sublimation
radius. We place a narrow ring at 0.23 AU with a width
of 0.035 AU (i.e., ∆r/r ∼0.15) to mimic the pile-up ef-
fect. For simplicity, we adopt the same grain parameters
as in the drag-in disk for this pile-up ring.
We then adjust the amount of material in each of
these components in the combined SED: the asteroid belt
(same as before, just different normalization), the drag-
in disk and the pile-up ring. To make sure these SED
parameters produce a vertical optical depth distribution
that is consistent with the theoretical expectation, we
also compute the corresponding optical depth in the SED
7
Figure 5. Optical depth in our SED models in comparison with
the low collisional rate case (long dash line) and high collisional
rate case (dash-dot line) calculations. The model with a pile-up
ring (upper panel of Figure 6) is shown as the dash line, and the
model without a pile-up ring (bottom panel of Figure 6) is shown
as the solid line. Both SED models have the same asteroid belt (at
∼13 AU) and hot ring (at ∼0.1 AU).
model as following:
σndr
2πr
amax(cid:90)
τ (r) =
= Σ(r)
Qabsπa2f (a)da,
(1)
sponding surface density distribution (i.e., Σ(r) =(cid:82) ndz
where n is the dust number density and Σ(r) is the corre-
amin
), σ is the cross section of a particle, i.e., 2πa2Qabs for
a grain radius of a and absorption coefficient Qabs, and
f (a) is the normalized grain size distribution. The inte-
gration limits are the grain size boundaries used in the
SED model; i.e., for the drag-in disk, the limits are 3 and
10 µm for astronomical silicates. We assume a constant
surface density in the P-R disk; therefore, the resultant
optical depth is flat for this part. Figure 5 shows the
model optical depth distribution for two of the fits in
comparison with the theoretical values for the low and
high collisional rate cases. As shown in Figure 5, the
material required in the SED fit for the P-R disk is lower
than the maximum amount from the low collision case,
but higher than the collision-dominated case at distances
far way from the asteroid belt.
The final component in our model is the narrow ring
that gives rise to the K-band excess. According to the
modeling in Rieke et al. (2015), nano grains can be photo-
electrically charged and magnetically trapped inside the
dust sublimation radius (∼0.23 AU for silicates or ∼0.1
AU for carbon-like grains around Fomalhaut) under a
nominal condition (a dipole magnetic field of 1 G around
an A-type star), and the gyroradius (i.e., the inner edge
of the trapped particles) could get as close as 0.05 AU.
For simplicity, we adopt a constant-surface-density ring
from 0.1 to 0.15 AU, composed of amorphous carbon
nano grains with sizes ranging from 0.01 µm (10 nm)
to 0.05 µm (50 nm) in an a−4 size distribution. The
exact grain composition and size parameters do not have
a huge impact on the SED model as long as they produce
a Rayleigh-Jeans-like spectrum from ∼2 -- 10 µm that best
fits the K-band interferometric measurement.
There are four individual components in our final SED
model: the asteroid belt, the P-R disk, the pile-up ring,
and the magnetically trapped hot ring. As a result,
many combinations of the scaling for each of the com-
ponents (i.e., the total dust mass) give satisfactory fits
0.11.010.0radius (AU)10−610−510−410−3τ at 0.55 µmSED models Low Collision CaseHigh Collision Case8
Su et al.
Figure 7. Null levels measured by the KIN in 2008 (squares with
error bars) from Mennesson et al. (2013) in comparison with the
model null levels (lines). The upper panel shows the short (58 m)
baseline result, while the lower one is for the long (76 m) baseline
result. Our best-fit models are shown as the solid line (without a
pile-up ring, i.e., the bottom panel of Figure 6) and the dotted line
(with a pile-up ring, i.e., the top panel of Figure 6). The null levels
from the Lebreton et al. (2013) model are also shown (the thick
dashed line) for comparison.
especially, the model with a pile-up ring simply gives
slightly higher null levels than those of the model with-
out. Our SED model for the pile-up ring (the upper
panel of Figure 6) represents the maximum amount of
emission in the model that is consistent with the overall
SED. This model gives an enhancement factor of ∼10
in the optical depth, which is the most extreme case
found in the study of Kobayashi et al. (2011) while a
typical enhancement around an A-type star for silicate
grains is ∼3. Since the expected emission from the maxi-
mum pile-up ring is very minimal compared to the nano-
grain hot ring and the P-R disk (see the upper panel of
Figure 6), the available measurements cannot determine
whether such a pile-up ring exists. (3) The wavelength-
dependent behavior in the KIN short baseline measure-
ments (the upper panel of Figure 7) is difficult to match
with our simple axi-symmetric model. Future observa-
tions from LBTI (Hinz et al. 2012) can help to confirm
this wavelength-dependent behavior, and provide com-
plementary spatial constraints that are crucial to distin-
guish between degenerate KIN models (e.g., Defr`ere et
al. 2015).
Table 2 summarizes the SED parameters in each of
the components. Note that the current data do not put
a strong constraint on the existence of a pile-up ring near
the silicate sublimation radius; we list its SED parame-
ters just for the sake of completeness. With or without
the ring due to pile-up, our model appears to give a sat-
isfactory fit to the KIN data (Figure 7), as does also the
Lebreton model (details see Section 5.3).
5. DISCUSSION
5.1. Does the Derived Dust Mass Make Sense?
Is the amount of dust in the asteroid belt too much for
a belt undergoing normal collisional evolution? A simple
way to answer this question is to compare the observed
fractional luminosity with the maximum fractional lumi-
nosity expected from collisional evolution models, fmax
Figure 6. SED models to fit all the excess measurements in Fo-
malhaut's inner 20 AU region. The data points are the same as in
Figure 3. The total emission (thick dashed line) is the sum of four
different components: a 13-AU narrow asteroid belt (solid line),
a flat disk due to P-R drag (triple-dot dashed line), and a hot
ring composed of magnetically trapped nano grains (dotted line).
The top panel shows the model that includes a narrow ring due to
the pile-up effect near the inner edge of the P-R disk (dot-dashed
line) and the bottom panel shows the same model but without the
pile-up ring.
in the overall SED, and Figure 6 shows two of them.
Our final constraint comes from the spatial information
in the KIN measurements. Each of the good-fit SED
models produces a slightly different spatial flux distri-
bution at the KIN wavelengths (8 -- 13 µm), i.e., different
wavelength-dependent null levels (the fraction of trans-
mitted flux detected by KIN) at different baselines (res-
olutions). We first simulate the high-resolution, face-on
model images at 7 -- 14 µm using the best-fit SED models.
Assuming that the inner dust structures have the same
inclination and position angles as the cold belt (i=66◦
and P.A.=156◦), the model images are then inclined and
rotated accordingly. Using these model images, we then
compute the expected null levels on the given dates and
baselines of the KIN observations (Table 2 of Mennesson
et al. 2013), and compare them with the observed levels.
Figure 7 shows examples of the comparison.
By changing the combination of individual components
and the associated SED parameters (like a Gaussian ring
vs. a flat disk), we explore the parameter space that the
KIN measurements are sensitive to. We find that (1)
an extended dust component in the ∼AU region (the
proposed P-R disk) is needed to explain the KIN mea-
surements, i.e., a hot nano-grain ring at ∼0.1 AU is not
enough to produce the KIN signals, corroborating the
finding in Mennesson et al. (2013). (2) The KIN mea-
surements are not sensitive to the changes in the dust
structure within ∼0.3 AU (i.e., the hot and pile-up rings);
1101001000wavelength (µm)100101102103flux density (mJy)Total EmissionAsteriod BeltPR Drag−in DiskPile−Up RingHot Ring1101001000wavelength (µm)100101102103flux density (mJy)Total EmissionAsteriod BeltPR Drag−in DiskHot RingParameters in the four SED components
Table 2
q
amax
[µm]
1000 −3.65 Gaussian Ring
−5.5
10
−5.5 Gaussian Ring
10
−4.0
0.05
Flat Disk
Flat Disk
[M⊕]
1.5×10−5
1.4×10−7
7.1×10−10
8.0×10−11
9
fd
Component
Dust Type
amin
[µm]
Density Type† Rin/Rp Rout/Rw Dust Mass
[AU]
[AU]
Asteroid Belt
P-R Disk
Pile-up Ring‡
Hot Ring
8.8×10−6
1.3×10−6
5.2×10−6
9.7×10−4
We use two types of surface density distribution: (1) a Gaussian ring characterized as the peak radius (Rp) and the
Silicates
Silicates
Silicates
Carbon
13
0.23
0.23
0.1
0.035
0.15
2
10
0.01
5
3
3
†
width (Rw), and (2) a flat disk characterized as the inner (Rin) and outer (Rout) radii.
‡
Current data do not put a strong constraint on the existence of a pile-up ring.
(i.e., equation (21) from Wyatt et al. (2007)). After scal-
ing for the mass and luminosity of Fomalhaut, we find
fmax is ∼1×10−5 for a belt at 10 AU at an age of 450
Myr. The observed fractional luminosity is ∼0.8×10−5
derived from our model, implying the amount of dust
observed in the belt does not violate the collisional evo-
lution models5. The ∼10 AU ice-line belt can be an
in-situ planetesimal belt.
The optical depth distribution in our simple P-R disk
model (shown in Figure 5) is between the low and high
collision cases expected from theoretical calculations ex-
cept for the region right interior of the inner belt. We
also used a similar gradual curve in that part of the P-R
disk, but it made no difference in the output SED and
resultant KIN nulls. This is not a surprise given that this
region is outside the field of view of KIN, and SED mod-
els are not sensitive to changes in small spatial scale. The
total derived dust mass in the P-R disk is 1.4×10−7M⊕,
which is ∼100 times less than the derived dust mass in
the asteroid belt (1.5×10−7M⊕), suggesting the ice-line
belt can sustain the mass needed in the P-R disk. Overall
the derived dust masses in the asteroid belt and the P-R
disk agree with the collisional and dynamical evolution.
For the hot ring, our derived mass is 8×10−11M⊕.
This value is very similar to the mass (∼10−10M⊕) de-
rived from other studies (Absil et al. 2009; Lebreton et
al. 2013; Rieke et al. 2015).
In other words, the mass
required to explain the VLTI K-band excess is on the
order of ∼10−10M⊕. The question remains whether the
P-R disk can supply enough material to form a nano-
grain hot ring. We can estimate the mass flow near
the inner boundary of the P-R disk (without a pile-
up ring) as
c2 Qprτ (r) in the low collisional
rate case (equation (9) from van Lieshout et al. (2014)).
With L∗ = 17.4 L(cid:12), Qpr = 1 for ∼µm-size grains, and
τ ∼ 4×10−6 from our model, the mass flow rate is
∼2×10−12M⊕yr−1 at 0.23 AU (where P-R grains sub-
In comparison, the maximum mass flow rate
limate).
from P-R drag at 0.23 AU is 3×10−11M⊕yr−1, using van
Lieshout's equation (11). These rates and the mass re-
quired in the hot ring suggest a typical resident time
of ∼5 -- 50 years for the nano grains. The lifetime of the
trapped nano grains is estimated to be ∼months to years
depending on the grain sizes (Rieke et al. 2015). The es-
timated resident time is longer than the lifetime of nano
grains if they are only supplied from the P-R disk. This
implies that the P-R disk in our model can be a source
MPR(r) = L∗
5 The criterion is usually set when fd >100 fmax (Wyatt et al.
2007).
to supply the nano grains; however, other mechanisms
like star-grazing comets might be needed to supplement
this mechanism.
5.2. How Sensitive are These Models to the Adopted
Grain Compositions?
The choice of the grain composition in the hot ring
does not really affect the outcome of the model as long
as the output SED is similar to Rayleigh-Jeans or steeper
(details see Rieke et al. (2015)). We adopt amorphous
carbon grains because they are a common material that
has a high sublimation temperature. Other oxides like
FeO can also work as well (see the discussion in Su et al.
2013 and Rieke et al. 2015).
We adopt only one composition, the silicate-like grains
(astronomical silicates), in both the asteroid belt and
the P-R disk because they are the most common mate-
rial found in interplanetary dust particles (IDPs). As a
result, the model SEDs show weak bumps in the 10 and
20 µm silicate features because the smallest size in the
model is ∼3 µm, just small enough to give rise to such
features. A mixture with modest amounts of organic ma-
terial would also produce a similar result with additional
adjustments (like the dust sublimation radius) on the ex-
act grain properties. The weak silicate features will be
washed out when using a mixture with other featureless
material. The exact numbers of the best-fit grain prop-
erties and density distribution are expected to vary, de-
pending on the adopted grain compositions. The overall
behavior in the link between the dust-producing asteroid
belt and the interior, low-density, P-R disk remains the
same.
5.3. Comparison Between Our and Lebreton's Models
Lebreton et al. (2013) have presented the most com-
plete and detailed model of the Fomalhaut inner debris
disk to date. Since their model is different from ours, it
is important to clarify what data are included in their
fit and the main difference between the two models. The
primary data used to constrain Lebreton model come
from the spatial constraints from the KIN 8 -- 13 µm mea-
surements, the VLTI K-band excess, and the broadband
photometry at 24 and 70 µm. They propose two distinct
dust populations in the Fomalhaut inner disk: (1) a very
narrow ring confined at 0.1 -- 0.3 AU from the star and
composed of unbound, very small 0.01 -- 0.5 µm carbon-
like grains, and (2) a disk peaked at ∼2 AU with a r−1
density distribution outward and composed of silicates
and carbon grains in a very steep power-law size distri-
bution (index of −4.8 to −4.1) with a minimum cutoff
10
of ∼ 2 -- 3 µm. Figure 3 also shows their model SED.
This model clearly over-predicts the flux in the 15 -- 30
µm range (the IRS data were not used as a constraint),
and the 24 µm excess is ∼1 Jy (see their Figure 5), which
is higher than the unresolved excess of 0.64 ±0.13 Jy es-
timated in Section 4.1. This is likely due to their lower
photospheric estimate, ∼2.78 Jy at 24 µm (see their Fig-
ure 5).
Several scenarios were discussed and explored in Le-
breton et al. (2013) to explain the sources of dust in the
∼2-AU region, which include: 1.) in-situ dust produc-
tion, i.e., a planetesimal belt at 2 AU6; 2.) P-R drag-in
grains from the cold (∼50 K) outer belt; and 3.) plan-
etesimal delivery scattered inward by multiple planets
inside the cold belt. They ruled out the in-situ dust pro-
duction since the observed fractional luminosity is a few
orders of magnitude higher than what a 2-AU belt in col-
lisional equilibrium can produce at the age of Fomalhaut
(450 Myr). In comparison, the derived dust properties
in our ice-line belt at ∼10 AU region are consistent with
collisional evolution of an in-situ planetesimal belt as dis-
cussion in Section 5.1. The P-R drag from the outer belt
was also ruled out because the amount of drag-in grains is
inadequate to explain the dust level at the 2 AU region.
They conclude that inward scattering of planetesimals
from the outer belt by a chain of tightly packed planets
is marginally adequate.
The innermost, hot ring poses more challenges as in
other previous studies because its SED requires emission
dominated by very small grains that are unstable against
photon pressure blowout. Lebreton et al. (2013) include
a detailed treatment for the dust sublimation process to
produce the very small grains in the hot ring and derive a
replenishing rate of ∼8×10−8M⊕yr−1 to sustain the hot
ring. Conventional mechanisms to supply such a flow
fall far short of this value; for example Lebreton et al.
find that, P-R drag from the 2 AU belt is inadequate by
nearly four orders of magnitude. Even invoking an evap-
orating planet came up short, although the yield from
such an event is very uncertain. Bonsor et al. (2014)
developed the comet scattering hypothesis further and
found that, with some very closely-packed inner planets
and an outermost planet migrating into the cold plan-
etesimal belt, it could supply a rate slightly more than
10−9M⊕yr−1, still an order of magnitude too small. In
addition, this hypothesis depends on an ad hoc arrange-
ment of planets and is implausible to be operating equally
well around the many other stars with similar very hot
excess emission.
To mitigate these problems, Lebreton et al. (2013) pro-
posed that the very small grains are trapped against es-
cape by gas in the hot ring, but this solution has signifi-
cant problems. Such a gas disk with normal (hydrogen-
rich) abundances would be easily detected in optical
emission lines (it would be analogous to the disks around
Ae/Be stars -- see, e.g., Mendigut´ıa et al. 2015). Instead,
Lebreton et al. (2013) suggest that the gas is the residue
from the sublimation of dust, and show that a total mass
of 5×10−3M⊕ would then suffice. In this case, emission
in the CII I58 µm line would be expected (Zagorovsky et
6 The typical dust temperatures in this region are ∼400 -- 450 K,
which is very different from the dust temperature we refer as the
ice-line belt.
Su et al.
al. 2010). Cataldi et al. (2015) have used Herschel/PACS
to place an upper limit on the emission from this line
from the entire Fomalhaut ring system, and show that
the signal from the spaxal centered on the star (see their
Figure 1) is even less than from the rest of the cold
ring. Assuming the gas coincident with the very hot
dust would be at a similar temperature, the upper limit
is an order of magnitude lower than the required mass in
the gas disk hypothesized by Lebreton et al. (2013).
In our model, the excess detected by the long-baseline
KIN observation arises from the P-R drag-in grains from
the ∼10-AU ice-line belt. These drag-in grains then sub-
limate as they get close to the sublimation radius (0.23
AU for silicate-like grains) and become the charged nano
grains trapped by the the weak magnetic field of the star
as proposed by Rieke et al. (2015), forming the hot ring.
The lifetime of these nano grains (months to years) then
suggests a replenishing rate of ∼10−10M⊕yr−1. As dis-
cussed in Section 5.1, the P-R mass inflow rates are in
the range of 10−12− 10−11M⊕yr−1 from the ice-line belt;
like star-grazing
an additional source of nano-grains,
comets, may be needed to supplement the P-R inflow
from the ice-line belt. This star-grazing comet delivery is
different from the cometary delivery scenario presented
in Bonsor et al. (2014) where a very specific planetary
configuration is required to sustain the high influx rate
of comets. There will be a wide range of planetary con-
figurations that both meet the scattering criteria as eval-
uated by Bonsor & Wyatt (2012) and allow to scatter a
few star-grazing comets per year into the inner part of
the Fomalhaut planetary system (including the tightly
pack Bonsor et al. (2014) configuration).
5.4. The Ice-line Asteroid Belt -- A Natural Source for
the Dust Interior to the Belt
The prominent cold belt cannot supply enough mate-
rial to the inner 2-AU region as discussed by Lebreton et
al. (2013). With our new parameters for the ice-line belt
(dust mass of 1.5×10−5M⊕ and collisional time scale of
2×104 yr), the required mass flow rate, ∼10−9M⊕yr−1,
is still too high to be sustained by the outer 140 AU
cold belt due to PR drag (the maximum mass flow is
∼10−12M⊕yr−1, equation (11) from van Lieshout et al.
2014). Unlike Eri (MacGregor et al. 2015), the 7 mm
ATCA observation of Fomalhaut (Ricci et al. 2012) does
not reveal any convincing excess emission above the pho-
tosphere. Also ionized winds from A-type stars like Fo-
malhaut are extremely weak (Babel 1995). Therefore, it
is unlikely that stellar-wind drag can aid the P-R drag
significantly. An in-situ ice-line belt, as we have demon-
strated, is the best alternative to supply the interior dust
detected by infrared interferometry.
The amount of interior dust that can be delivered from
a planetesimal belt may even be overstated by the P-R
drag estimates. In our Solar System, the dust level inte-
rior to the Kuiper belt is thought to be constant into ∼10
AU, where most of the particles are ejected by Saturn
and Jupiter (Moro-Mart´ın & Malhotra 2005), and only
a very small amount of the P-R drag-in dust can reach
Earth's vicinity (Liou & Zook 1999; Kuchner & Stark
2010; Vitense et al. 2010). These model predictions are
consistent with the measurement by the Student Dust
Counter (Szalay et al. 2013, 2015) on board the New
Horizons Mission. In addition to the denseness of a belt
(collision rate), giant planets located interior to a plan-
etesimal belt can also reduce the amount of dust interior
to the belt. To verify whether the amount of material
due to P-R drag interior to a planetesimal belt is lower
than the expected value (i.e., a sign of an additional de-
pletion mechanism like the presence of giant planet(s)),
a detailed collisional cascade calculation is needed. How-
ever, such a model requires detailed information about
the planetesimal belt (i.e., location and mass) that is
currently lacking.
6. CONCLUSION
We report an ALMA cycle 1, 870 µm observation cen-
tered at the star position of the Fomalhaut planetary sys-
tem. We detect a point source, consistent with the bare
stellar photosphere, and no extended structures that are
brighter than 3σ in the central 15×15 AU region. We
evaluate all available measurements to constrain the in-
frared excess arising from dust in the inner 20 AU region
in Fomalhaut, and conclude that a dust-producing plan-
etesimal belt at the ice sublimation temperature (i.e., an
asteroid-belt analog) is the most likely origin for the in-
frared excess longward at ∼15 µm. The location of the
ice-line belt is estimated to be at ∼8 -- 15 AU using SED
models with nominal parameters for a narrow belt with
a 3σ upper limit of total flux less than 1.3 mJy at 870
µm. Assuming the inner belt has the same orientation
as the outer one (inclination and position angles of 66◦
and 156◦), but centered at the star position, we detect
a few 1 -- 2σ blobs along the expected disk circumference
of a 13-AU belt. Although our SED model suggests such
a 13-AU belt might be detected by our cycle 1 data at
1 -- 2σ levels, the putative 13-AU belt needs future confir-
mation.
We further propose a new coherent model to explain
the interferometric hot excesses by connecting the pro-
posed asteroid belt to the dust structures inside of it.
We suggest that a small amount of material from the
ice-line belt can spiral inward under the influence of P-
R drag, forming an interior, extended disk composed of
dust grains that have β values closer to 0.5. The inner
boundary of the P-R disk is set by the dust sublimation
radius of the dominant material like silicates (∼0.23 AU
around Fomalhaut when grains reach ∼1300 K). We also
consider a possible pile-up ring near the silicate sublima-
tion radius as proposed by Kobayashi et al. (2009) with
an enhancement factor of a few in the optical depth. We
show that the required optical depth of such an interior
disk (and a pile-up ring) is lower than the maximum al-
lowable amount through the P-R drag under the theoret-
ical calculations. Finally, the sublimation of these drag-
in silicate grains is likely to produce nano carbon-like
or FeO-like particles that should be blown out by radia-
tion pressure very quickly under a normal circumstance.
However, a weak (∼1 G) magnetic field from a fast rotat-
ing A-type star like Fomalhaut can efficiently trap these
photoelectrically charged nano grains forming a hot ring.
The location of the hot ring depends on the magnetic
field strength and the grain sublimation temperature. As
modeled by Rieke et al. (2015), a ring at ∼0.1 AU and
composed of nano, very refractory grains around Foma-
lhaut can explain the K-band excess detected by VLTI.
The required resident time of the nano grains sustained
11
entirely from the P-R disk is somewhat longer than the
collisional lifetime of these nano grains; therefore, addi-
tional sources like a few star-grazing comets are needed
to explain the hot ring. Combining all these components
(the asteroid belt, the P-R disk with and without a pile-
up ring, and the hot nano-grain ring), we can simultane-
ously obtain good fits to the excess SED and satisfy the
spatial constraints set by the KIN measurements.
Studying the inner zones in debris disks requires high
angular resolution that can only be provided in scat-
tered light by ground-based high-contrast imaging fa-
cilities and/or HST, and by ALMA in thermal emis-
sion. However, the scattered light study of the inner
debris structures around A-type stars like Fomalhaut is
challenging because the inner debris is intrinsically faint
due to lower mass (compared to the outer debris) and
lack of efficient scatterers of small grains (typical radi-
ation blowout size for Fomalhaut is ∼3 µm). Although
future LBTI and JWST mid-infrared observations can
trace ∼µm-size grains in the inner zone in great detail,
their structures are much more influenced by radiation
forces (radiation pressure and P-R drag). Only ALMA
can resolve the detailed structure of the asteroid belt
by tracing the mm-size grains that show the imprint of
extrasolar terrestrial planets through dynamical interac-
tions. Knowing the detailed properties of the asteroid
belt not only will help us better understand the processes
creating tenuous dust structures interior of the belt, but
also provides insights into extrasolar terrestrial planets
that we cannot probe otherwise. Deeper observations
with the full ALMA array have the potential to reveal
directly the asteroid belt inferred from our SED models.
This paper made use of the following ALMA data:
ADS/JAO.ALMA 2012.0.00238.S (PI: S.P.L.). ALMA is
a partnership of ESO (representing its member states),
NSF (USA), and NINS (Japan), together with NRC
(Canada) and NSC and ASIAA (Taiwan), in coopera-
tion with the Republic of Chile. The Joint ALMA Ob-
servatory is operated by ESO, AUI/NRAO, and NAOJ.
K.Y.L.S. is grateful for funding from NASA's ADAP pro-
gram (grant number NNX11AF73G). Support for G.H.R.
is provided by NASA through contract 1255094 and
1256424 issued by JPL/Caltech to the University of Ari-
zona. S.P.L thanks the support of the Ministry of Science
and Technology (MoST) of Taiwan with Grants NSC
98-2112-M-007-007-MY3, NSC 101-2119-M-007-004 and
MoST 102-2119-M-007-004-MY3.
REFERENCES
Absil, O., Mennesson, B., Le Bouquin, J.-B., et al. 2009, ApJ,
704, 150
Acke, B., Min, M., Dominik, C., et al. 2012, A&A, 540, AA125
Alekseeva, G. A., Arkharov, A. A., Galkin, V. D., Hagen-Thorn,
E. I., et al. 1996, Baltic Ast., 5, 603
Babel, J. 1995, A&A, 301, 823
Backman, D., Marengo, M., Stapelfeldt, K., et al. 2009, ApJ, 690,
1522
ApJ, 775, 55
Ballering, N. P., Rieke, G. H., Su, K. Y. L., & Montiel, E. 2013,
Boley, A. C., Payne, M. J., Corder, S., et al. 2012, ApJ, 750, LL21
Bonsor, A., & Wyatt, M. C. 2012, MNRAS, 420, 2990
Bonsor, A., Raymond, S. N., Augereau, J.-C., & Ormel, C. W.
2014, MNRAS, 441, 2380
Cataldi, G., Brandeker, A., Olofsson, G., et al. 2015, A&A, 574,
L1
12
Su et al.
Chiang, E., Kite, E., Kalas, P., Graham, J. R., & Clampin, M.
Mendigut´ıa, I., de Wit, W. J., Oudmaijer, R. D., et al. 2015,
MNRAS, 453, 2126
Mennesson, B., Absil, O., Lebreton, J., et al. 2013, ApJ, 763, 119
Mermilliod, J. C. 1991, "Catalog of Homogeneous Means in the
UBV System,", online at Vizier
Min, M., Dullemond, C. P., Kama, M., & Dominik, C. 2011,
Icarus, 212, 416
730, L29
Morales, F. Y., Rieke, G. H., Werner, M. W., et al. 2011, ApJ,
Moro-Mart´ın, A., & Malhotra, R. 2005, ApJ, 633, 1150
Oteo, I., Zwaan, M. A., Ivison, R. J., Smail, I., & Biggs, A. D.
2015, arXiv:1508.05099
Quillen, A. C. 2006, MNRAS, 372, L14
Ricci, L., Testi, L., Maddison, S. T., & Wilner, D. J. 2012, A&A,
Rieke, G. H.; Blaylock, M.; Decin, L.; Engelbracht, C. et al. 2008,
Rieke, G. H., Gaspar, A., & Ballering, N. P. 2015,
539, L6
AJ, 135, 2245
arXiv:1511.04998
Stapelfeldt, K. R., Holmes, E. K., Chen, C., et al. 2004, ApJS,
Su, K. Y. L., Rieke, G. H., Misselt, K. A., et al. 2005, ApJ, 628,
Su, K. Y. L., Rieke, G. H., Stapelfeldt, K. R., et al. 2009, ApJ,
Su, K. Y. L., Rieke, G. H., Malhotra, R., et al. 2013, ApJ, 763,
154, 458
487
118
705, 314
Su, K. Y. L., & Rieke, G. H. 2014, IAU Symposium, 299, 318
Szalay, J. R., Piquette, M., & Hor´anyi, M. 2013, Earth, Planets,
and Space, 65, 1145
Szalay, J., Piquette, M., & Horanyi, M. 2015, Lunar and
Planetary Science Conference, 46, 1701
van Lieshout, R., Dominik, C., Kama, M., & Min, M. 2014, A&A,
571, A51
Vernazza, J. E., Avrett, E. H., & Loeser, R. 1976,ApJS, 30, 1
Vitense, C., Krivov, A. V., Lohne, T. 2010, A&A, 520, A32
Wiegert, J., Liseau, R., Th´ebault, P., Olofsson, G. et a.
Wyatt, M. C. 2005, A&A, 433, 1007
Wyatt, M. C., Smith, R., Greaves, J. S., et al. 2007, ApJ, 658, 569
Wyatt, M. C., Clarke, C. J., & Booth, M. 2011, Celestial
Mechanics and Dynamical Astronomy, 111, 1
Zagorovsky, K., Brandeker, A., & Wu, Y. 2010, ApJ, 720, 923
PASP, 119, 994
1981, A&AS, 45, 5
74
2009, ApJ, 693, 734
Chen, C. H., Mittal, T., Kuchner, M., et al. 2014, ApJS, 211, 25
Defr`ere, D., Hinz, P. M., Skemer, A. J., et al. 2015, ApJ, 799, 42
Dermott, S. F., Jayaraman, S., Xu, Y. L., Gustafson, B. A. S., &
Liou, J. C. 1994, Nature, 369, 719
Engelbracht, C. W., Blaylock, M., Su, K. Y. L., et al. 2007,
Engels, D., Sherwood, W. A., Wamsteker, W., & Schultz, G. V.
G´asp´ar, A., Psaltis, D., Rieke, G. H., & Ozel, F. 2012, ApJ, 754,
G´asp´ar, A., Rieke, G. H., & Balog, Z. 2013, ApJ, 768, 25
Greaves, J. S., Sibthorpe, B., Acke, B., et al. 2014, ApJ, 791, L11
Hinz, P., Arbo, P., Bailey, V., et al. 2012, Proc. SPIE, 8445,
Holland, W. S., Greaves, J. S., Zuckerman, B., et al. 1998,
84450U
Nature, 392, 788
1345
ApJ, 775, 56
Johnson, H. L. 1965, Comm. Lun. Plan. Lab., 3, 73
Kalas, P., Graham, J. R., & Clampin, M. 2005, Nature, 435, 1067
Kalas, P., Graham, J. R., Chiang, E., et al. 2008, Science, 322,
Kalas, P., Graham, J. R., Fitzgerald, M. P., & Clampin, M. 2013,
Kennedy, G. M., & Piette, A. 2015, MNRAS, 449, 2304
Kennedy, G. M., & Wyatt, M. C. 2014, MNRAS, 444, 3164
Kobayashi, H., Watanabe, S.-i., Kimura, H., & Yamamoto, T.
2009, Icarus, 201, 395
Kobayashi, H., Kimura, H., Watanabe, S.-i., Yamamoto, T.,
Muller, S. 2011, Earth, Planets, and Space, 63, 1067
Kuchner, M. J., & Stark, C. C. 2010, AJ, 140, 1007
Laor, A., & Draine, B. T. 1993, ApJ, 402, 441
Lebreton, J., van Lieshout, R., Augereau, J.-C., et al. 2013, A&A,
Liou, J.-C., & Zook, H. A. 1999, AJ, 118, 580
Liseau, R.; Montesinos, B.; Olofsson, G.; Bryden, G., et al. 2013,
555, A146
A&AL, 549, 7
A&AL, 573, 4
Loukitcheva, M. A., & Nagnibeda, V. G. 2000, "The Solar Cycle
and Terrestrial Climate," ESA Publications, 463, 363
MacGregor, M. A., Wilner, D. J., Andrews, S. M., Lestrade,
J.-F., & Maddison, S. 2015, ApJ, 809, 47
Marengo, M., Stapelfeldt, K., Werner, M. W., et al. 2009, ApJ,
Matthews, B., Kennedy, G., Sibthorpe, B., et al. 2014, ApJ, 780,
700, 1647
97
McMullin, J. P., Waters, B., Schiebel, D., Young, W., & Golap,
K. 2007, Astronomical Data Analysis Software and Systems
XVI, 376, 127
Liseau, R.; Vlemmings, W.; Bayo, A.; Bertone, E., et al. 2015,
2014,A&A, 563, 102
|
1312.1585 | 1 | 1312 | 2013-12-05T15:43:04 | Feasibility Studies for the Detection of $O_2$ in an Earth-like Exoplanet | [
"astro-ph.EP"
] | We present the results of simulations on the detectability of $O_2$ in the atmosphere of Earth twins around nearby low mass stars using high resolution transmission spectroscopy. We explore such detectability with each of the three upcoming Extremely Large Telescopes (ELTs), i.e. GMT, TMT and E-ELT, and high resolution spectrographs, assuming such instruments will be available in all ELTs. With these simulations we extend previous studies by taking into account atmospheric refraction in the transmission spectrum of the exo-Earth and observational white and red noise contributions. Our studies reveal that the number of transits necessary to detect the $O_2$ in the atmosphere of an Earth twin around M-dwarfs is by far higher than the number of transits estimated by Snellen et al. (2013). In addition, our simulations show that, when accounting for typical noise levels associated to observations in the optical and near-infrared, the $O_2$ A-band at 760 nm is more favorable to detect the exoplanetary signal than the $O_2$ band at 1268 nm for all the spectral types, except M9V. We conclude that, unless unpredicted instrumental limitations arise, the implementation of pre-slit optics such as image slicers appear to be key to significantly improve the yield of this particular science case. However, even in the most optimistic cases, we conclude that the detection of $O_2$ in the atmosphere of an Earth twin will be only feasible with the ELTs if the planet is orbiting a bright close-by (d $\le$ 8 pc) M-dwarf with a spectral type later than M3. | astro-ph.EP | astro-ph |
Feasibility Studies for the Detection of O2 in an Earth-like
Exoplanet
Florian Rodler
Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA
Institut de Cienci`es de l'Espai (CSIC-IEEC), Campus UAB, Fac. Ciencies, C5 p2, 08193 Barcelona, Spain
and
Harvard-Smithsonian Center for Astrophysics, Cambridge, MA 02138, USA
Mercedes L´opez-Morales
ABSTRACT
We present the results of simulations on the detectability of O2 in the atmosphere of Earth
twins around nearby low mass stars using high resolution transmission spectroscopy. We explore
such detectability with each of the three upcoming Extremely Large Telescopes (ELTs), i.e. GMT,
TMT and E-ELT, and high resolution spectrographs, assuming such instruments will be available
in all ELTs. With these simulations we extend previous studies by taking into account atmo-
spheric refraction in the transmission spectrum of the exo-Earth and observational white and red
noise contributions. Our studies reveal that the number of transits necessary to detect the O2 in
the atmosphere of an Earth twin around M-dwarfs is by far higher than the number of transits
estimated by Snellen et al. (2013). In addition, our simulations show that, when accounting for
typical noise levels associated to observations in the optical and near-infrared, the O2 A-band at
760 nm is more favorable to detect the exoplanetary signal than the O2 band at 1268 nm for all
the spectral types, except M9V. We conclude that, unless unpredicted instrumental limitations
arise, the implementation of pre-slit optics such as image slicers appear to be key to significantly
improve the yield of this particular science case. However, even in the most optimistic cases, we
conclude that the detection of O2 in the atmosphere of an Earth twin will be only feasible with
the ELTs if the planet is orbiting a bright close-by (d ≤ 8 pc) M-dwarf with a spectral type later
than M3.
Subject headings: astrobiology -- atmospheric effects -- planetary systems -- techniques: spectroscopy
1.
Introduction
The path to finding exoplanets either inhabited
or amenable for life as we know it on Earth in-
cludes the detection in their atmospheres of chem-
ical compounds such as H2O, CO2, CH4, and
O3, known as biomarkers (Schindler et al. 2000,
Pavlov et al. 2000, Kaltenegger & Traub 2007,
Traub et al. 2008). Another important biomarker
is gaseous oxygen, O2, detected by Sagan et al.
(1993) while analyzing the spectrum of Earth ob-
served by the Galileo probe. In that spectrum O2
appeared to be, together with CH4 in strong ther-
modynamical disequilibrium, the strongest indica-
tor of life on Earth. O2 has since drawn significant
attention because it provides a strong indication of
the presence of oxygen-producing forms of life and
because it has several spectral absorption bands at
visible and near-infrared wavelengths, which can
be detected using ground-based telescopes. The
most prominent of those bands are the so-called
O2 A-band around 760 nm, and the O2 band at
1268 nm, with the A-band being the strongest
with 55 strong lines.
1
Several studies have now investigated the de-
tectability of O2 for an Earth-like exoplanet in
the habitable zone of its host star. The first of
those studies was published by Schneider (1994)
who analytically calculated the detectability of the
O2 A-band by estimating the photometric depth
of the absorption produced by the full band, which
has a width of ∼ 3 nm. Computing the ratio of
the projected surface areas of the planet's atmo-
sphere and the disk of the star during a transit,
this study concluded that it was possible to detect
the O2 A-band with a 2.4 meter space-bound tele-
scope, assuming the planet orbits a solar type star
at a distance of 2 parsecs or a 0.3 R⊙ star at 10
parsecs.
Webb & Wormleaton (2001) refined the Schnei-
der (1994) analysis by using empirical observations
of the Earth's O2 A-band to simulate the transmis-
sion spectrum of the exoplanet and by superim-
posing the resultant spectrum with the spectrum
of the host star. Their analysis concluded that it
is possible to detect the planetary O2 A-band for
host stars with R < 0.3 R⊙, although detectabil-
ity with current 8-meter class telescopes would be
limited to M dwarfs with mv ∼ 10. This study
also proposed that a way to confirm that the O2
lines originated in the atmosphere of the exoplanet
is that they should be offset from the Earth's own
O2 lines by an amount equal to the peculiar ve-
locity of the host star1. However, they did not
include the telluric O2 A-band in their analysis.
A few years later, Kaltenegger & Traub (2009)
generated simulations using the Earth's atmo-
sphere as proxy to estimate the detectability of
biomarkers during the transit of an Earth twin
around a Sun-like star and also around M-dwarfs,
for a 6.5-meter telescope in space. Their main
conclusion was that the signal-to-noise ratio (S/N)
of a single transit would not be enough to detect
any spectral features in the atmosphere of an exo-
Earth orbiting in the habitable zone of those stars.
Instead, multiple transits would be needed, with
an estimate of 6 to 192 transits per year for M0V to
M9V stars. They also concluded that the spectral
features of different molecules in the atmosphere
1References in the literature trace this technique back to
an article by Lowell (1905), where he proposes that oxy-
gen lines in the atmosphere of Mars could be detected by
their Doppler shift with respect to the corresponding tel-
luric lines.
of an exo-Earth around an M star will be easier to
detect in the infrared than in the optical, because
the stellar flux peaks in the infrared.
Just recently, Snellen et al.
(2013) extended
previous studies by including the effect of the tel-
luric O2 A-band and applying the latest tech-
niques to detect the planet's Doppler shift from
the ground (e.g. Brogi et al. 2012, Rodler et al.
2012). Their simulations show that it will be pos-
sible to detect the O2 A-band in the atmosphere
of an Earth twin if a spectrograph with a spec-
tral resolving power of R = 100, 000 is installed
on the next generation of Extremely Large Tele-
scopes (ELTs). In particular, they concluded that
the 39-meter E-ELT, will be able to detect the O2
A-band transmission signal of a planet around an
M5V star by combining the observations of a few
dozen transits. In addition, they remarked that in
the case of stars smaller than M7V the O2 might
be easier to detect in the O2band at 1268 nm since
those stars will be brighter at near-infrared wave-
lengths than in the visible.
In this paper we further refine the Snellen et
al. (2013) simulations by 1) accounting for atmo-
spheric refraction effects in the exoplanet's trans-
mission spectrum and 2) including the effect of
random and correlated (atmospheric and instru-
mental) observational noise. In addition, we ex-
tend previous studies by exploring the expected
performance of instruments for the three upcom-
ing ELTs, exploring a range of spectral resolutions,
and performing simulations for a range of stellar
peculiar velocities and spectral types. Finally, we
explore in detail the feasibility of the observations
both in O2 A-band and the O2 band at 1268 nm.
In sections 2 and 3 we describe the implementa-
tion and results of our simulations. A discussion of
those results and the optimal observational set-up
is given in section 4.
2. Simulations
2.1.
Ingredients
The first step to perform these O2 detection
feasibility studies is to generate a set of hypotheti-
cal exo-Earth transiting planet configurations and
observational setups. The parameters necessary
for the simulations are 1) a model transmission
spectrum of the Earth-like planet (which we as-
sume to be an Earth twin, including atmospheric
2
conditions and chemistry), 2) a model transmis-
sion spectrum of Earth, 3) a model of the spectral
emission of the star, including assumptions about
the relative velocity of the star with respect to
Earth, 4) a telescope+instrument setup configu-
ration, and 5) observational noise models.
2.1.1. Exoplanet Model Transmission Spectrum
The model atmospheric transmission spectrum
of the exo-Earth was calculated adopting a line-
by-line radiative transfer model (LBLRTM) code
based of the FASCODE algorithm (Clough et al.
2005)2. As molecular database for oxygen models
we adopted HITRAN (Rothman et al. 2009).
In those calculations we accounted for the ef-
fect of refraction in the transmission spectrum of
the exo-Earth. As described in Garcia-Munoz et
al. (2012), the transmission spectrum of an Earth
twin observed transiting its host star will differ
from the Earth's transmission spectrum because of
refraction. Refraction is the effect by which light
crossing the planet's atmosphere near its surface
gets deflected by an angle that depends on the
properties of the atmosphere itself, the size of the
star, and the orbital distance of the planet. This
effect causes the atmospheric layers close to the
planet's surface to be invisible for a remote ob-
server. As a consequence, any chemical species at
heights lower than that limit can not be observed.
For species distributed over a large range of atmo-
spheric heights, as is the case of O2, the depth of
their absorption lines will be altered. For e. g., in
the particular case of the Earth-Sun system atmo-
spheric layers at heights lower than 12-14 km will
be invisible to a remote observer seeing the system
transit. In the case of smaller stars, i.e. M-dwarfs,
where both the radius of the star and the orbital
distance of an Earth-like planet in the habitable
zone of the system are smaller, only atmospheric
altitudes lower than about 5 km appear invisible.
In this work we focus on Earth twins orbiting
around M-dwarf stars (see section 2.1.3). There-
fore, to simulate this refraction effect we only in-
tegrate the model transmission spectrum of the
exo-Earth at atmospheric layers between 5 km to
2LBLRTM
code
source
available
http://rtweb.aer.com/lblrtm description.html
fortran
platforms.
and manuals
is
runs
available
on
various
as
and
code
are
source
The
under
3
85 km from the surface. 85 km marks the end of
the Earth's mesosphere, with over 99% of the at-
mospheric mass contained below that atmospheric
height (Lutgens & Tarbuck 1995). Therefore, the
contribution of the atmospheric layers above that
height to the transmission spectrum of the planet
can be considered minuscule.
The resulting model transmission spectrum of
the exo-Earth around the O2 A-band as well as
the O2 band at 1268 nm are illustrated in the top
panels Figures 1 and 2.
2.1.2. Telluric Spectrum
The transmission spectrum of Earth (telluric
spectrum) was calculated following the procedure
outlined by Seifahrt et al. (2010). Like in the pre-
vious section, this procedure uses the LBLRTM
code developed by Clough (2005) and the trans-
mission molecular absorption database HITRAN
containing the 42 most prominent molecules and
isotopes present in the atmosphere of Earth. As
an input to LBLRTM, we adopted a model of
the Earth's atmosphere that contains meteorolog-
ical information for temperature and pressure as a
function of height for an average night at a given
observatory. In our case we retrieved the weather
information from the Global Data Assimilation
System (GDAS). GDAS models are available in 3
hour intervals for any location around the globe3.
By employing LBLRTM, we then calculated ab-
sorption and emission spectra for a given path
through the atmosphere, i.e. for a given airmass.
In our model we have assumed that the observa-
tions take place at a ground-based observatory at
an altitude of about 2500 meters above sea level,
and we have accounted for a changing airmass val-
ues during the course of the night within the range
X = 1.1 to 2. The transmission spectrum around
the O2 A-band and the O2 band at 1268 nm for an
airmass of 1.3 is illustrated as example in the mid-
dle panels of Figures 1 and 2. We note that a few
H2O absorption lines appear in the telluric spec-
trum around 1290 nm. These lines partly over-
lap with the O2 lines and constitute an additional
source of noise. In our simulations, we have con-
sidered the depth of these lines to be constant for
simplicity reasons.
In fact, however, the depths
of the water absorption lines may strongly vary
3GDAS webpage: http://ready.arl.noaa.gov/READYamet.php
i
i
n
o
s
s
m
s
n
a
r
T
i
i
n
o
s
s
m
s
n
a
r
T
x
u
F
l
1.0
0.8
0.6
0.4
1.0
0.5
0.0
1.0
0.5
0.0
760
762
764
766
Wavelength (nm)
768
770
Fig. 1. -- Top panel: transmission spectrum of
the atmosphere of an Earth-like planet around
760 nm. Middle panel: telluric spectrum of our
atmosphere for a zenith distance of 30◦ (airmass
X = 1.3). Bottom: PHOENIX model spectrum
of an M4V star with a surface temperature of
Teff = 3000 K, log g = 4.5 dex and solar abun-
dance. All spectra are shown at a spectral resolu-
tion of R = 100, 000.
i
i
n
o
s
s
m
s
n
a
r
T
i
i
n
o
s
s
m
s
n
a
r
T
x
u
F
l
1.0
0.8
0.6
0.4
1.0
0.5
0.0
1.0
0.5
0.0
1250
1260
1270
1280
1290
Wavelength (nm)
Fig. 2. -- Same as in Fig. 1, but for the wavelength
regime around 1268 nm.
4
during the course of a night (e.g. Rodler et al.
2012), an effect that should be implemented in fu-
ture simulations.
2.1.3. Stellar Spectra
To model the emission from the star, we fo-
cus on M-type stars with spectral types M1V --
M9V. The reason for this, as already explained
by Kaltenegger & Traub (2009) and Snellen et al.
(2013), is that the planetary transit depths are in-
versely proportional to the squared radius of the
star and therefore, the transmission spectrum sig-
nal of an exo-Earth will be easier to detect around
a star with smaller radius. For a star like the Sun
(G2V), the ratio of the areas between the stellar
disk and the atmosphere of an exo-Earth's ring is
of the order of 5× 105, while for an M1V star that
ratio is ∼ 1.25×105, and for an M9V star the ratio
drops to ∼ 4×103.
We adopted stellar model spectra for effective
temperatures and surface pressures corresponding
to M1V -- M9V stars using the new library of high-
resolution spectra based on the PHOENIX code
and presented in Husser et al. (2013). This new
library has been generated using a new equation
of state, which improves the treatment of chemical
equilibrium in stellar atmospheres, and updated
atomic and molecular line lists. These new mod-
els provide a better match to observed spectra of
M dwarfs than the latest PHOENIX (Brott et al.
2005) and ATLAS9 (Castelli et al. 2003) models.
For simplicity, we assumed solar abundances in all
the models. As an example, we show the model
spectrum generated for an M4V star with effective
temperature Teff = 3000 K, gravity log g = 4.5 dex
and solar abundances in the bottom panels of Fig-
ures 1 and 2.
2.1.4. Telescope+Instrument Setup
There are three Extremely Large Telescopes
(ELTs) with diameters 24 meters or larger ex-
pected to start operations in the next decade. The
smallest of those ELTs is the Giant Magellan Tele-
scope (GMT; www.gmto.org) with an effective di-
ameter of 24.5 meters. One of the first light instru-
ments on the GMT will be G-CLEF (Szentgyor-
gyi et al. 2012), which will be a high-resolution
spectrograph with resolution modes between R
= 25,000 -- 120,000 and spectral coverage between
350 -- 950 nm, suitable to study the O2 A-band.
The second largest of the ELTs is the Thirty
Meter Telescope (TMT; www.tmt.org). Among
the TMT's expected suite of instruments the best
suited to detect O2 in the atmosphere of an Earth-
like planet is the High Resolution Optical Spec-
trometer (HROS: Crampton et al. 2008), listed
among the telescope's first decade instruments.
HROS will provide a resolution of R = 50 000
for a 1 arcsec slit, or R > 90 000 with an image
slicer, and spectral coverage between ∼ 310 -- 1000
nm. An image slicer is a device that allows to feed
a large fraction of the stellar flux into a narrow slit,
thereby avoiding slit losses and at the same time
allowing to employ high spectral resolving powers
(see e.g. Dekker et al. 2003).
The largest of the ELTs will be the European
Extremely Large Telescope (E-ELT), with a diam-
eter of 39-meters. In the case of the E-ELT, there
are two spectrographs with resolution R > 100 000
being studied on Phase A4, CODEX (Pasquini
et al. 2010) and SIMPLE (Origlia et al. 2010).
CODEX only covers wavelengths up to ∼ 710 nm,
while the wavelength coverage of SIMPLE starts
at ∼ 800 nm. Therefore none of these instru-
ments include the O2 A-band at 760 nm. SIM-
PLE, on the other hand, will cover the O2 band
at 1268 nm. A third high-resolution spectrograph
called HIRES, is now being proposed as one of the
first light instruments for the E-ELT, although the
specifications for this instrument are not defined
yet.
2.1.5. Noise Models
The last of the simulation inputs is noise. The
noise in an observed spectrum can be divided into
random (white) and correlated (red) noise. While
the distribution of white noise is Poissonian and
can be reduced by increasing the S/N of the spec-
trum, the red noise is not random and cannot be
reduced by increasing the S/N. However, the red
noise can be related to instrumental and atmo-
spheric effects, such as e.g. variations of airmass
or changes of the position of the spectrum on the
detector. Analysis techniques have improved in
recent years to levels where it is now possible to
significantly reduce the amount of red noise in the
data.
In our simulations we have added both sources
of noise. White noise is simulated using a random
distribution function with a maximum amplitude
set by the S/N of the particular model (see sec-
tion 2.2).
In the case of red noise, its pattern
can contain a series of frequencies that depend on
the specific cause of the noise for a given dataset.
Therefore, we have simulated red noise using the
approach described in (Kasdin 1995), by assuming
a functional form of 1/f, where f is the frequency.
The amplitude of the red noise is set to a fraction
of the white noise in an specific model. For our
simulations we consider red noise contributions of
0 to 100% of the white noise. Typical red noise
levels of ground based data remain of the order
of 20% of the white noise in the optical and 50-
100% of the white noise in the near-infrared (see
e.g. Pont et al. 2006, Rogers et al. 2009).
2.2. Generation of Model Spectra
Using the inputs described above we generate
model spectra, C, using a expression of the form
C = (cid:16)a (1 + ǫ−1)−1T(cid:17) ⊗ G ,
(1)
where
a = (1 + v⋆c−1) S + (cid:0)1 + (v⋆ + vpl)c−1(cid:1) ǫ−1 P .
(2)
In these equations, c denotes the speed of light,
S is the spectrum of the host star, P is the trans-
mission spectrum of the planetary atmosphere,
and T is the telluric spectrum of Earth for a given
airmass. G is the instrumental profile (IP) of
the spectrograph. In our simulations, we assume
the IP to be a Gaussian function that degrades
the spectral resolution of the combined spectrum
to a chosen value. The remaining parameters in
the equations are v⋆, the relative stellar velocity
with respect to Earth (i.e. systemic velocity and
barycentric velocity), vpl, the instantaneous radial
velocity of the planet with respect to its host star
(which is close to 0 km s−1 at a primary transit),
and ǫ, the area ratio between the stellar disk and
the atmospheric ring of the planet.
Once we have computed the spectrum C, we
interpolate it onto a pixel grid, which simulates
a given instrumental layout. Since the exact pa-
rameters of most instruments described in section
2.1.4 are not yet fully defined, we adopt two dif-
4http://www.eso.org/sci/facilities/eelt/instrumentation/phaseA.html
5
ferent pixel grid scales; one with a velocity res-
olution of 0.75 km s−1 pixel−1 and the other
1.2 km s−1 pixel−1. The 0.75 km s−1 pixel−1 grid
scale corresponds to the planned scale of G-CLEF
for resolutions R > 25, 000 (Szentgyorgyi et al.
2012), while the 1.2 km s−1 pixel−1 grid scale cor-
responds to UVES (Dekker et al. 2000), which
is currently mounted at the 8.2 Very Large Tele-
scopes (VLT) of the European Southern Observa-
tory (ESO).
The next step is to adjust the flux of the model
spectra based on the expected flux of the exo-
Earth's host star that is received on Earth. Given
that the potential transiting exo-Earth host stars
will be at a range of possible distances from Earth,
and therefore will have different apparent magni-
tudes, we consider a range of brightnesses for stars
of a given spectral type.
(Snellen et al. 2013)
and (Kaltenegger et al. 2009) computed the most
likely apparent magnitude of a nearby transiting
exo-Earth host star based on the solar neighbour-
hood mass function and transit probability of an
exo-Earth and they provided detectability estima-
tions that are based on that number (see e. g. Sec-
tion 3 in Snellen et al. 2013). Here we adopt
a different approach and instead of making any
assumption about the apparent magnitude of the
transiting exo-Earth host star, we assume that
such star can be at any distance within 20 pc.
Based on the star's distance, we adjust its appar-
ent magnitude between 1 and 20 pc 5. The effec-
tive temperature and radius of the stars adopted
for each spectral type are listed in Table 1. The
table also shows the values of the area ratio be-
tween the stellar ring and the atmospheric ring of
the planet, ǫ, the I-band and J-band magnitudes of
each star at 10 pc, the orbital period of a planet in
the middle of the habitable zone of the star based
on the definition by Kasting (1993), and the du-
ration of the transits of such planet.
The flux that we finally measure from a star of a
given magnitude will depend on the telescope and
instrument throughput. Therefore, we estimated
5We start our models at 1 pc based in the assumption that
there is no other star closer to the Sun than Proxima Cen-
tauri at 1.33 pc. The 20 pc upper limit is chosen because of
the faintness of M-dwarfs. In the remaining paper we only
consider distances up to 11 pc since exo-Earths around stars
further than that distance take too long to observe (see sec-
tions 3 and 4).
6
those fluxes using online Exposure Time Calcula-
tors (ETC) for each instrument, whenever avail-
able. For all ETCs, we assumed a median seeing
of 0.8" and airmass values ranging from 1.1 to 2.0.
G-CLEF already has an ETC6, while for SIMPLE,
we adopted the throughput value estimates listed
on the project webpage7. The instrumental layout
of HROS has not been published yet; we therefore
assumed that HROS will be similar to G-CLEF,
with a factor of 1.5 shorter integration times due
to the larger diameter of the TMT with respect
to the GMT. For a spectrograph with a UVES-
like layout, we used the UVES ETC8 and scaled
the results accordingly to the diameter ratio be-
tween the ELT and the VLT with 8.2 m. For this
instrument, we calculated the S/Ns for different
slit widths, resulting in different spectral resolv-
ing powers. In addition, we also determined the
S/N when employing the UVES image slicer #3,
which has a resolution of R = 110, 000 (Dekker et
al. 2003).
With the ETCs we estimated the exposure time
needed to attain the maximum number of counts
for each star's apparent magnitude, but at the
same time avoiding reaching typical non-linearity
regimes of 36 000 ADU and 12 000 ADU (in the
brightest pixel on the chip) in a single exposure,
respectively for the CCDs and near-infrared de-
tectors. We furthermore accounted for the inte-
gration time and the dead time between two ex-
posures to estimate the observations duty cycle
for each star. While for CCDs we assumed a dead
time of 15 seconds and a read-out noise of 5 elec-
trons per exposure, we accounted for a dead time
of 10 seconds and a read-out noise of 10 electrons
for infrared detectors. The maximum single expo-
sure time was chosen to be 600 seconds, since at
longer exposures the radial velocity shift of the re-
mote planet would smear out the absorption lines
in its atmospheric spectrum.
Combining the duty cycle of each star with the
flux collected per integration we obtained the total
number of photons collected during the duration of
one transit (see last column in Table 1). The duty
cycles for each star, for each instrument, are sum-
ETC
version
1.0;
6G-CLEF
http://alerce.astro.puc.cl/gclef.html
7SIMPLE http://simple.bo.astro.it/
8ETC
http://www.eso.org/observing
version
3.2.13;
Website:
marized in Tables 2 and 3. The remaining columns
in these tables are explained in section 3.4.
Finally, we added Poisson noise and red noise
to each model based on their flux levels (i.e. S/N),
following the description in section 2.1.5. Before
adding the noise, we scaled each spectrum in such
a way that the actual S/N per spectral pixel cor-
responded to the square root of the flux at that
pixel. This approach ensures that regions of tel-
luric lines as well as stellar absorption lines have a
higher noise and a lower S/N than the continuum.
2.3. Data sets and their analysis
We generated data sets simulating transmission
spectra of planetary transits using eqs. 1 and 2
with the following range of parameters, v⋆ = −150
to 150 km s−1, ǫ = 4 × 103 − 5.2 × 105, depending
on the spectral type of the host star, and spectral
resolutions of R = 60, 000 − 110, 000, as well as a
set of S/N levels including red noise contributions
from 0 to 100% of the white noise level. In each
data set, we simulated the observations during a
night. Hence, each data set consisted of a series
of spectra, which differed from each other by the
airmass and the instantaneous radial velocity shift
of the planetary transmission spectrum (vp) with
respect to its host star during transit. Addition-
ally, we factored in the effects of the change of the
barycentric velocity of the star plus its planet dur-
ing the course of a night, which is of the order of a
few times 100 m s−1. Furthermore, to rule out se-
lection effects coming from the adopted pixel grid,
we also varied the zero-point of the pixel grid up
to the average velocity span of one pixel. The
duration of the datasets corresponds to twice the
duration of the transits listed in the last column
of Table 1.
To analyze each transit dataset in search for
the extraterrestrial signal of O2, we followed the
approach outlined in Rodler et al. (2012, 2013)
and removed both a model telluric spectrum and
a model stellar spectrum from each synthetic spec-
trum, thereby calculating a residual spectrum
which -- ideally -- only contained the transmis-
sion spectrum of the planet, plus noise. We then
carried out cross-correlations between the resid-
ual spectrum and the transmission model spec-
trum of the exo-Earth, and determined the cross-
correlation function. For all spectra of each data
set, we then summed up the cross-correlation func-
7
tions in the rest frame of the planet and deter-
mined the radial velocity value that yielded the
highest correlation (candidate feature).
To determine the number of false positives (i.e.
the confidence level of the candidate feature), we
employed a bootstrap randomization method (e.g.
Rodler et al. 2013). Random values of the or-
bital phases were assigned to the observed spec-
tra, thereby creating N different data sets (N =
100, 000 in our analyses). Any signal present in
the original data was then scrambled in these ar-
tificial data sets. For all these randomized data
sets, we re-ran the data analysis in the given pa-
rameter search ranges and located the candidate
feature with its maximum value.
The confidence level of the candidate features
was estimated to be ≈ 1 − FAP = 1 − m/N ,
where FAP is the false alarm probability and m
is the number of the best fit models having a peak
value larger or equal than the peak value of the
cross-correlation function found in the original,
unscrambled data sets.
A detection is considered significant when the
signal is at least 3σ, which corresponds to a FAP
of ≤ 0.0027.
3. Results
Using the simulations described above, we have
investigated a series of questions pertaining the
successful ground-based detection of O2 in the at-
mosphere of an Earth twin around a nearby star.
Those questions include the optimal spectral res-
olution of the observations, the limitations im-
posed by the relative radial velocity of the star
and Earth, the limitations imposed by the depth
of the telluric O2 lines, and the number of transits
we need to observe to achieve at least a 3σ detec-
tion in the presence of white and red noise. The
answers to those questions are provided in Figures
3-8 and Tables 2-4, and are also described in detail
in the text below.
The simulations were carried out for G-CLEF
mounted on the 24m GMT, for HROS on the TMT
assuming an instrument configuration similar to
G-CLEF, and for three instruments mounted on
the E-ELT; the near-infrared spectrograph SIM-
PLE and two hypothetical high-resolution spec-
trographs in the visual, a G-CLEF-like instrument
and UVES-like instrument.
Host Star and Habitable Zone Planet Parameters
Table 1
spectral type
Teff (K)
R (R⊙)
ǫ
MI (mag) MJ (mag)
P (d)
transit duration (h)
G2V
M1V
M2V
M3V
M4V
M5V
M6V
M7V
M8V
M9V
5800
3600
3400
3250
3100
2800
2600
2500
2400
2300
1
0.49
0.44
0.39
0.26
0.20
0.15
0.12
0.10
0.08
520 000
125 000
101 000
80 000
35 000
21 000
12 000
7 500
6 000
4 000
4.1
7.7
8.3
8.8
10.0
11.2
12.4
13.6
13.9
14.7
3.6
6.4
6.5
7.1
7.9
8.6
10.1
10.7
11
11.6
365.2
43
33
27
16
9.5
6.0
4.1
3.3
1.9
13.1
4.0
3.4
3.0
2.1
1.5
1.1
0.78
0.69
0.43
Simulations for E-ELT spectrographs in the visual†
Table 2
sp. type
obs. time (h)
duty cycle
transits
time (years)
obs. time (h)
UVES-like design
G-CLEF-like design
duty cycle
transits
time (yrs)
G2V
M1V
M2V
M3V
M4V
M5V
M6V
M7V
M8V
M9V
840
110
98
88
43
46
43
35
34
29
0.05
0.58
0.69
0.79
0.92
0.97
0.98
0.98
0.98
0.98
65
28
29
29
21
31
39
45
48
66
≥ 65
∼ 29
∼ 23
∼ 19
∼ 8
∼ 7
∼ 6
∼ 5
∼ 4
∼ 3
429
67
62
57
29
31
30
24
25
22
0.08
0.71
0.80
0.87
0.95
0.98
0.98
0.98
0.98
0.98
34
17
19
19
14
21
27
31
36
50
≥ 33
∼ 18
∼ 15
∼ 13
∼ 6
∼ 5
∼ 4
∼ 3
∼ 3
∼ 2
†For a velocity span of 1.2 km s−1 pixel−1 (UVES-like design, left) and 0.75 km s−1 pixel−1 (G-CLEF-like design, right).
Numbers are given for 5 pc distance.
Simulations for G-CLEF @ GMT and HROS @ TMT
Table 3
sp. type
obs. time (h)
duty cycle
transits
time (years)
obs. time (h)
duty cycle
transits
time (yrs)
G-CLEF @ GMT
HROS @ TMT
G2V
M1V
M2V
M3V
M4V
5V
M6V
M7V
M8V
M9V
470
133
133
130
70
79
75
61
69
67
0.18
0.86
0.91
0.94
0.98
0.98
0.98
0.98
0.98
0.98
37
33
40
44
34
53
68
78
100
154
≥ 37
∼ 35
∼ 31
∼ 28
∼ 14
∼ 12
∼ 10
∼ 8
∼ 8
∼ 7
448
97
95
91
48
53
51
41
46
45
0.13
0.81
0.87
0.92
0.97
0.98
0.98
0.98
0.98
0.98
35
24
28
30
23
36
46
53
67
104
≥ 35
∼ 26
∼ 23
∼ 20
∼ 9
∼ 9
∼ 7
∼ 6
∼ 6
∼ 5
For a distance of 5 pc
8
Table 4
Simulations for SIMPLE @ E-ELT
Sp. type
obs. time (h)
duty cycle
transits
time (yrs)
M3V
M4V
M5V
M6V
M7V
M8V
M9V
713
173
99
38
21
18
10
0.12
0.41
0.69
0.75
0.84
0.87
0.91
238
82
66
35
27
26
24
≥ 100
∼ 33
∼ 15
∼ 5
∼ 3
∼ 2
∼ 1
For a distance of 5 pc
3.1. Spectral Resolution
We used the UVES ETC to investigate the ef-
ficiency of observations with an UVES-like instru-
ment at different spectral resolutions. We carried
out simulations for resolutions between 60,000 and
110,000, rescaling the S/N of the contimuun of
each spectrum to the aperture of the E-ELT. Fig-
ure 3 shows the result of those tests for the partic-
ular case of an Earth twin orbiting around an M4V
star at a distance of 5 pc. In this case, the opti-
mal spectral resolution appears to be R = 80, 000,
which corresponds to the best trade-off between
slit losses and the spectral smearing of the exo-
Earth's signal. The plot also shows, for compar-
ison, the result of the same test when using an
image slicer -- in this particular case we simulated
image slicer #3 described in Dekker et al. (2003).
When using the image slicer, which is only used
at resolution R = 110, 000, the efficiency of the
observations improves by a factor of two with re-
spect to the efficiency of the instrument at that
same resolution when using only a slit. We note
that this result represents the ideal case, i.e. we
did not account for possible changes of the instru-
mental profile induced by an image slicer, which
might hamper the data analysis. The tests also
show that the use of the image slicer improves the
overall efficiency of the observations, where only
about 21 transits are necessary to achieve a 3σ
detection of the exo-Earth's atmosphere, while 25
transits are necessary to achieve the same preci-
sion level when just observing at a spectral reso-
lution of R = 80, 000 without an image slicer.
The result of these tests are instrument-specific
and similar tests will have to be done for other in-
struments to determine the optimal resolution for
this type of science once their final specifications
are defined. However, we can argue that employ-
ing pre-slit optics, such as image slicers capable of
concentrating a larger fraction of the stellar flux
into the slit, can significantly reduce the number
of transits needed for a successful detection, for all
instruments.
3.2. Relative Radial Velocity
The detectability of an Earth twin depends,
among other things, on how many of its O2 lines
can be observed. Our atmosphere produces the
same absorption features (telluric lines), at the
same wavelength positions (cf. Figs. 1 and 2), and
in addition, the lines have some intrinsic width.
Therefore, the Earth twin spectrum will be only
detectable when its lines are Doppler shifted with
respect to the telluric lines by certain amounts.
To investigate this effect we generated a set of
simulations in which the spectrum of the Earth
twin was shifted with respect to the spectrum of
Earth by values of the relative velocity between
the host star and Earth between -150 km s−1 and
+150 km s−1, and determined the fraction of lines
that would appear blended in each case. The re-
sults of those simulations are illustrated in Figure
4 for both the O2 A-band and the O2 band at 1268
nm.
In the case of the O2 A-band, we find that
most lines will appear blended as long as the rel-
ative velocities are < ± 15 km s−1. Line blend-
ing becomes significant again (with about 50% of
the lines blended) at relative velocities of about ±
50 km s−1, coinciding with the average separation
between individual lines in the O2 A-band spec-
trum. The optimal relative velocity regimes (with
10% or less of the lines blended), appear to be -15
to -30 km s−1 and 15 to 30 km s−1.
9
In the case of the O2 band at 1268 nm, which
are narrower than the A-band lines, we find that
relative radial velocities in the regimes around -
85 to -65 km s−1, -10 to 10 km s−1, and 65 to
85 km s−1 are also significantly blended. Stars
with any other relative velocity value can be ob-
served with the guarantee that no more than about
20% of the O2 lines will be blended.
3.3. Telluric Line Depths
Another effect affecting the detectability of an
Earth twin is noise introduced by telluric con-
tamination. In other words, telluric lines block a
large fraction of the host star's light, causing the
S/N of its spectrum to drop significantly around
those wavelengths. In the case of the O2 band at
1268 nm, telluric contamination around 1268 nm
causes the S/N ratio to drop to 0.3 with respect
to the S/N ratio in the continuum. In the case of
the O2 A-band, telluric contamination around 760
nm fully blocks all incoming light at some wave-
lengths (see middle panels in Figures 1 and 2).
The best strategy to avoid this source of noise dur-
ing the data analysis process is to skip wavelength
regions where the S/N of the observed spectrum
drops below a certain value. Therefore, we de-
fine a cut-off parameter with respect to the S/N
of the spectral continuum for which lines (or por-
tions of the spectrum) to keep during the analysis.
The optimum value of this cut-off parameter is a
trade-off between keeping a large enough number
of absorption features from the planet and avoid-
ing noisy regions due to strong telluric contami-
nation. Adopting a non-optimal cut-off parameter
will result in strongly suppressing the exo-Earth's
signal and might lead to a non-detection (i.e. to a
detection with confidence level less then 3σ).
To determine the value of this cut-off parame-
ter we analyzed different datasets by masking out
different wavelength regions in which the telluric
absorption lines dropped below a certain amount,
e.g.
a cut-off parameter of 0.5 corresponds to
masking out all the wavelength regions for which
the S/N drops by 50% with respect to that of the
continuum. The result of these tests is shown in
Figure 5, where we represent the FAP of the exo-
Earth's detection as a function of the cut-off pa-
rameter in S/N units (i.e. a value of 1 for the
cut-off parameter means that even the continuum
is masked out). The figure indicates that the best
10
without IS
with IS
s
t
i
s
n
a
r
t
f
o
r
e
b
m
u
n
35
30
25
20
60
80
70
100
spectral resolution (R/1000)
90
110
Fig. 3. -- Simulation results for the number of
transit observations necessary to obtain a 3σ de-
tection of O2 around 760 nm in the atmosphere of
an Earth twin orbiting a M4V star at 5pc. The
instrumental setup is an UVES-like instrument
mounted on the 39-m E-ELT. The open circles
show the result of the simulations for spectral reso-
lutions between 60,000 and 110,000. The filled cir-
cle shows the result of the same simulations when
an image slicer at a resolution of 110,000 is used.
These simulations assume a relative radial veloc-
ity for the star of 20 km s−1 with respect to Earth
and no red noise.
-150
1
-100
-50
0
50
100
150
O2 A-band
s
d
n
e
b
l
e
n
i
l
f
o
n
o
i
t
c
a
r
f
s
d
n
e
b
l
e
n
i
l
f
o
n
o
i
t
c
a
r
f
0.8
0.6
0.4
0.2
0
1
0.8
0.6
0.4
0.2
O2 B-band
0
-150
-100
-50
0
50
100
150
relative radial velocity [km/s]
Fig. 4. -- Plot showing the normalised fraction of
line blends between the telluric spectrum and the
transmission spectrum of the Earth twin as a func-
tion of the relative radial velocity between the host
star and Earth. The upper panel corresponds to
the O2 A-band at 760 nm and the bottom panel
to the O2 band at 1268 nm. Based on this plot,
observations are best suited when the relative ve-
locity of the star is between ±(15 and 30) km s−1
in the case of the A-band, and for velocities out-
side the ranges ±(85 to 65) km s−1 and -10 to
10 km s−1 in the case of the O2 band at 1268 nm.
value of the cut-off parameter is approximately 0.4
times the S/N of the continuum.
3.4. Number of Transits
The last question we addressed is how many
transit observations are necessary to achieve at
least a 3σ detection of an Earth twin's atmosphere,
and what will be the time span needed to collect
those many transits. To estimate the number of
transits needed, we focused first on the 760 nm O2
A-band and assumed a transiting planetary sys-
tem with an Earth twin in its habitable zone, at a
distance of 5 parsecs from Earth and moving away
from us with a relative velocity of 20 km s−1. We
determined the duty cycle of the observations by
assuming an instrument dead time of 15 sec after
each exposure (see section 2.2). To calculate the
number of transits, we divided the observing time
by the transit duration of the planet as a func-
tion of stellar spectral type. The values for the
orbital periods and transit durations were taken
from Kaltenegger & Traub (2009) and are pro-
vided in Table 1. To estimate the overall time
span of the observations, we factored in the or-
bital period of the planet and assumed that only
every 9th transit can be observed, after account-
ing for object visibility from Earth, transit occur-
rence, and the relative radial velocity between the
target and Earth (see also Section 4). The results
are summarised in Tables 2 and 3.
Table 2 is split into two halves; the left-hand
side lists the results for a spectrograph with a
UVES-like layout (i.e.
spectral resolving power
of R = 110, 000, image slicer #3, velocity span
of 1.2 km s−1 pixel−1) installed on the E-ELT,
while the right-hand side shows the values de-
termined for a spectrograph with a design simi-
lar to G-CLEF (i.e. R = 100, 000, velocity span
of 0.75 km s−1 pixel−1), also on the E-ELT. G-
CLEF's design seems to be significantly more effi-
cient.
In this table, and in Table 3, we also show the
times needed for an Earth twin orbiting a G2V
star.
In this case, because of the very low duty
cycle and the very shallow transit depth, it would
take over 30 transits, and therefore more than 30
years to detect O2 with 3σ confidence. Once we
also factor in for the number of observable transits
from the ground, these observations would take
many decades. For M1V - M3V stars we estimate
11
the time span of the observations to be ∼ 13 -
19 years. For later type M-dwarfs, where the or-
bital periods of planets in their habitable zones are
shorter, this number drops significantly to about
6 years for M4V stars and about 2 years for M9V
stars. For a planet orbiting a M4V star, it would
require a minimum of 14 transits to detect O2 with
3σ confidence in its atmosphere.
In the case of
M9V stars, which are intrinsically fainter, it would
take at least 50 transits.
Figure 6 shows the number of transits required
for a 3σ detection with the two instrumental se-
tups in Table 2, as a function of distance and for
spectral types M3V through M7V. Planets around
an M3V star need more transits at very low dis-
tances due to a decrease in the duty cycle of the
observations, i.e. the stars become too bright and
the observing times need to be short to avoid sat-
uration. Therefore most of the observing time in
those cases is spent on instrument readouts. The
same applies for earlier type stars. The figure also
shows how for stars located at distances larger
than about 8-10 parsecs, the number of transits
necessary to achieve a 3σ detection becomes too
large.
Table 3 is similar to Table 2, but showing the
results for two other observational configurations;
the left-hand side of the table shows the results
of our simulations for G-CLEF mounted on the
GMT. The right-hand side shows the result of
the simulations for HROS mounted on the TMT.
For both spectrographs we assume a spectral res-
olution of R = 100, 000 and a velocity span of
0.75 km s−1 pixel−1. The top panel of Figure 7
shows the number of transits required for a 3σ de-
tection with G-CLEF on the GMT as a function
of distance and, as in Figure 6, for spectral types
M3V through M7V. As expected, in this case, the
number of transits necessary to achieve a detection
is larger, by approximately a factor of two, than
the number of transits necessary with an instru-
ment like G-CLEF in the larger aperture E-ELT
(also see bottom panel of Figure 6). The number
of transits in Figure 7 decrease by a factor of about
1.5 for the case of HROS mounted on the TMT.
Finally, Table 4 summarizes the results of our
simulations for the near-IR spectrograph SIMPLE
to be mounted on the E-ELT. In this case the sim-
ulations focus on the 1268 nm O2 band, while we
still assume a transiting planetary system with an
12
Earth twin in its habitable zone, at a distance of
5 parsecs from Earth and moving away from us
with a relative velocity of 20 km s−1. We adopted
a spectral resolution of R = 100, 000, a velocity
span of 1.2 km s−1 pixel−1 and a dead time be-
tween two subsequent exposures of 10 seconds. We
only carried out simulations for stars with spectral
types later than M3V, since the number of transits
necessary for an M3V star was already over 200.
The results on that table reveal that using the O2
band at 1268 nm to detect O2 in the atmosphere
of an Earth twin will require a significantly larger
amount of telescope time than observations of the
O2 A-band for all spectral types earlier than M7V.
The bottom panel of Figure 7 shows he number of
transits required or a 3σ detection with SIMPLE
as a function of distance and for spectral types
M5V through M9V. In this case, the number of
transits decreases steadily with increasing spectral
type.
3.4.1. The Effect of Red Noise
Tables 2, 3, and 4, as well as Figures 6 and 7
show the results for the ideal case, i.e. without any
contributions from correlated (red) noise. How-
ever, typical red noise levels are 20% and 50% of
the white noise level for data recorded in the vi-
sual and near-infrared, respectively (Pont et al.
2006, Rogers et al. 2009). The higher red noise
level in the near-infrared wavelength regime is due
telluric contamination and to a larger extent, due
to detector noise. To make the simulations more
realistic, we added different levels of red noise to
the simulated data sets and analysed them as de-
scribed before. Figure 8 shows the difference in
the amount of observing time required to achieve
a 3σ detection as the red noise levels increase. For
a red noise level of 20% of the white noise level, the
observing time required to attain a 3σ detection
rises by a factor of ∼ 1.4. This means that for a
red noise level of 20% of the white noise level, the
observing times as well as the number of transits
shown in Tables 2 and 3 and in Figure 6 and 7 (top
panel) needs to be multiplied by this value. This
effect becomes more severe for higher red noise
levels. e.g. for correlated noise levels of 50% and
100% of the white noise level, the time factor in-
creases by ∼ 2.6 and ∼ 9, respectively.
When comparing the results determined by
accounting for typical red noise levels (20% of
the white noise level in the visual, 50% of the
white noise level at near-infrared wavelengths),
we find that the lowest time span of observa-
tions is achieved with a G-CLEF-like spectrograph
mounted on the E-ELT for all M-dwarfs with spec-
tral types M1V-M8V. Observations in the O2 band
at 1268 nm are more efficient than in the visual
only for M9V stars (see Figure 9).
4. Discussion & Conclusions
We have carried out detailed feasibility studies
to probe molecular oxygen in the atmosphere of
an Earth-analogue orbiting M-dwarfs with future
instruments to be mounted on upcoming ELTs. In
our simulations, we factored in the stellar spectra
for different spectral types, the telluric spectrum
of the Earth, different instrumental settings of pos-
sible future instrumentation, refraction effects in
the Earth analog's atmosphere, and a variety of
noise models, including the effects of correlated
noise (red noise) in the data. When we compare
our work to previous, similar studies, these are the
most detailed simulations to date.
A result of our investigation is that a G-CLEF-
like instrument mounted on an E-ELT is the best
suited for the detection of O2 in the atmosphere
of an Earth-like planet orbiting M1V-M8V stars.
Only for planets orbiting M9V stars, observations
carried out in the near-infrared become more effi-
cient than in the visual. This is a result of lower
red noise levels at visible wavelengths, but the sit-
uation could change in the future if infrared red
noise levels can be improved to match noise levels
in the visible.
In that case, near-infrared obser-
vations will be more efficient for planets around
M7V stars and smaller.
Fig. 10 depicts the frequency of M1V to M9V
stars in the solar neighborhood9. M3 and M4
dwarfs are the most common sub-types, with more
than 25 stars of each spectral type within a dis-
tance of 10 pc. When combining this distribu-
tion with the results of our simulations, we con-
clude that very close-by M4V stars will be the
most likely candidates for the detection of oxy-
gen in the atmosphere of an Earth-like planet. In
just 40 hours of observing time (∼ 20 transits)
a 3σ detection of an object at 5 pc from Earth
9Data were
(http://www.recons.org).
retrieved from the RECONS database
13
could be attained with a G-CLEF-like instrument
mounted on the 39 m E-ELT. For this estimation,
we assumed a red noise level of 20% of the white
noise level. A typical orbital period of a planet lo-
cated in the habitable zone around an M4-dwarf is
16 days. Once the object visibility, orbital period
and the relative radial velocity are taken into ac-
count, we estimate that it would nevertheless take
a total time span of observations of ∼ 8 years to
accumulate sufficient data for a detection. This
number soars up to 175 hours (74 transits) and
a time span of roughly 30 years for an object at
a distance of 10 pc, which is the most probable
distance to find an Earth-analogue around such a
star (Kaltenegger & Traub 2009).
We compare our results to previous studies by
other groups. Snellen et al.
(2013) carried out
simulations with the goal to detect the O2 A-
band in the atmosphere of Earth-analogues orbit-
ing around M4V and M6V stars, with apparent
magnitudes I = 10 and I = 11.8, respectively.
According to their results, it takes about 4 to 15
transits to attain a 3σ detection of O2. They as-
sumed a line contrast ratio between stellar contin-
uum and the deepest absorption lines of 20, 000,
which corresponds to ǫ ≈ 35, 000 in the case of a
spectral resolution R = 100, 000, which causes a
line depth of 0.6 for the deepest lines with respect
to the stellar continuum (c.f. line depths in Fig. 1,
top panel). Furthermore, they assumed that the
transit duration was 1.4 hours for planets orbit-
ing those types of dwarfs. Adopting the values
for ǫ and the transit duration listed in Table 1,
our studies reveal that it takes a total of about
42 and 60 transits for a 3σ detection of oxygen
in the atmosphere of an Earth-like planet orbiting
an M4V and M6V stars, respectively. These re-
sults are for the ideal case, i.e. without red noise
and a G-CLEF-like spectrograph mounted on the
39m E-ELT. If we include a typical 20% red noise
contribution in these studies, the numbers rise to
60 and 84 transits (i.e. 12-25 years), respectively
for M4 and M6 dwarfs. Our results are far less
optimistic than those presented by Snellen et al.
(2013).
Ideally, the incoming flux from the planet host
star should find its way to the data analysis. To
record high-resolution spectra of a star, it is neces-
sary to carry out the observations through a nar-
row slit (or fiber). This, however, means that
only a fraction of the stellar light is used -- and
in case of bad seeing conditions, it might happen
that more than 60% of the incoming light is lost.
To avoid such an unfavorable situation, pre-slit
optics should be employed which enable to feed
more than 90% of the photons of the star into
the narrow slit. This can be achieved with im-
age slicers or multiple fibers, although in this last
case there is some concerns that transmission dif-
ferences between fibers might affect the results.
Such potential problems will have to be evaluated
before hand. Since the designs of most of the
high-resolution spectrographs to be mounted on
the ELTs haven't been fixed yet, we strongly rec-
ommend to include pre-slit optics to collect most
of the stellar light and feed it into the spectro-
graph. Based on our simulations in Figure 3, the
usage of pre-slit optics could improve the required
number of observations by a factor of two. In this
estimation, however, we have not accounted for
possible irregularities in the IP induced by an im-
age slicer, which might represent a further noise
source.
At this point, we want to emphasize that our
simulations represent a somewhat ideal case, i.e.
stable instrument configurations and perfectly fit-
ting models in the data analysis. Our simulations
need to be redone once the configuration details of
the spectrographs are known.
For the planning of the observations, we find
that a very important parameter to take into ac-
count is the relative radial velocity of the star with
respect to the Earth. That relative radial velocity
is the combination of the systemic radial velocity
of the star (vr) and the instantaneous barycentric
velocity vbary. Our atmosphere produces the same
line pattern that we search for in the atmosphere
of a remote planet. At some wavelengths, the tel-
luric contamination almost absorbs all the incom-
ing flux, leading to very low S/N ratios in those
wavelength regions, which need to be masked out
in the data analysis. It is therefore crucial to mini-
mize the numbers of absorption line blends by only
taking observation when the telluric spectrum and
the spectrum of the planet are Doppler shifted by
a certain amount. For observations in the visual,
we find that the optimum velocity range of the
planet host star with respect to the Earth is lim-
ited to two small velocity windows, ranging from
±(15 to 30) km s−1.
In Figure 11 we show the distribution of ab-
solute values of the barycentric radial velocities
for a sample of 176 M-dwarfs reported by Chubak
et al. (2012). Assuming that the distribution of
radial velocities in that sample is representative
of all nearby M-dwarfs, we can conclude that the
majority of them have barycentric radial velocities
between -15 and 15 km s−1. Therefore, what is the
amount of time of optimum visibility of those ob-
jects, i.e. when they show relative radial velocities
of ±(15 to 30) km s−1? To answer this question,
we need to account for the barycentric velocity of
the Earth, which causes a periodical radial veloc-
ity shift with a semi-amplitude Vbary of approxi-
mately 29.8 km s−1 for an object located close to
the ecliptic plane. However, the semi-amplitude of
the barycentric velocity is a function of the ecliptic
latitude β (i.e. the angular distance of the object
from the orbital plane):
Vbary = (29.8 cos
βπ
180
) km s−1.
(3)
For the vast majority of stars, the semi-amplitude
of the barycentric velocity Vbary is between 20 and
29.8 km s−1.
Figure 12 shows the amount of time of optimum
visibility of an object in units of the yearly observ-
ing time available at an ideal observing site. For
this estimation, we assumed that the object can
be observed at an airmass X ≤ 2 between 2 hours
and 8 hours (at opposition with Earth). Further-
more, for simplicity we assumed a circular orbit of
Earth and factored in the instantaneous value of
the barycentric velocity of the object. Figure 12
depicts that for semi-amplitudes of the barycentric
velocity of Vbary ≥ 20 km s−1, the atmospheric
lines from the planet would be detectable a frac-
tion of 0.2 to 0.4 of the time for the majority of
the M-dwarfs, which have instrinsic radial veloci-
ties between -15 and 15 km s−1.
To understand possible error sources in the data
analysis, we briefly review the main steps: We re-
move both the telluric spectrum as well as the stel-
lar spectrum from the data, and finally attempt
to retrieve the planetary signature from the resid-
ual spectrum via cross-correlation. The removal
of both spectra is an iterative process, which in-
volves the determination of the IP and the convo-
lution of the stellar + telluric model spectra with
the IP (the method is outlined in Rodler et al.
14
2013). While the telluric spectrum can be well re-
produced with theoretical model spectra, the mod-
eling of M-dwarf spectra is still in its infancy. In
principle, if the instrument is stable during one
night, it would be sufficient to take spectra of the
star for some hours before and after the transit,
subtract the telluric lines, then co-add these spec-
tra to form one high-S/N spectrum for that night,
which is then subtracted from the observed spec-
trum taken during the transit (cf. procedure used
by Rodler et al. 2008). However, this procedure
would drastically boost the need of observing time.
Detailed studies on the modeling of M-dwarf spec-
tra are clearly required.
Additionally, it is crucial to discard strong tel-
luric features in the data analysis. We want to
make future observers aware that regions which
show a S/N ratio less than about 0.4 times the
S/N ratio in the continuum of the spectrum should
be masked out. Adopting a wrong cut-off param-
eter causes the the weak planetary signal to be
suppressed.
Another possible error noise constitutes the
model spectrum of the planet atmosphere, which is
adopted in the cross-correlation. We are attempt-
ing to detect a very weak planetary signal of the
planetary atmosphere; the lack of some absorption
features in the theoretical model would suppress
the planetary signal in the cross-correlation func-
tion. To avoid this situation, a large numbers of
atmospheric models with different mixing ratios
and molecular configurations need to be probed.
We conclude that successful measurements of
O2 in Earth-like atmospheres with future ground-
based instrumentation will be most likely limited
to host stars of spectral types later than M3V,
which are located at distances less than 8 pc. Our
simulations demonstrate, that a 3σ discovery of
O2 in the atmosphere of such a planet would re-
quire even tenths of transits and several years of
observing time at ELTs.
We are grateful to the constructive comments of
the anonymous referee, who helped us to improve
the manuscript. FR acknowledges financial sup-
port from the Spanish Ministry of Economy and
Competitiveness (MINECO) and the "Fondo Eu-
ropeo de Desarrollo Regional" (FEDER) through
grant AYA2012-39612-C03-01. This research has
made use of NASA's Astrophysics Data System
15
Bibliographic Services.
REFERENCES
Brogi, M., Snellen, I. A. G., de Kok, R. J., Al-
brecht, S., Birkby, J., & de Mooij, E. J. W.
2012, Nature, 486, 502
Brott, I. & Hauschildt, P. H. 2005, in ESA Special
Publication, Vol. 576, The Three-Dimensional
Universe with Gaia, ed. C. Turon, K. S.
O'Flaherty, & M. A. C. Perryman, 565
Castelli, F. & Kurucz, R. L. 2003, in IAU Sym-
posium, Vol. 210, Modelling of Stellar Atmo-
spheres, ed. N. Piskunov, W. W. Weiss, & D. F.
Gray, 20P
Chubak, C., Marcy, G., Fischer, D. A., Howard,
A. W., Isaacson, H., Johnson, J. A., & Wright,
J. T. 2012, ArXiv e-prints
Clough, S. A. 2005, JQSRT, 91 233
Crampton, D., Simard, L., & Silva, D. 2008, in
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series, Vol. 7014,
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series
Dekker, H., D'Odorico, S., Kaufer, A., De-
labre, B., & Kotzlowski, H. 2000,
in Soci-
ety of Photo-Optical Instrumentation Engi-
neers (SPIE) Conference Series, Vol. 4008, So-
ciety of Photo-Optical Instrumentation Engi-
neers (SPIE) Conference Series, ed. M. Iye &
A. F. Moorwood, 534 -- 545
Dekker, H., Nissen, P. E., Kaufer, A., Primas, F.,
D'Odorico, S., & Hanuschik, R. W. 2003, in
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series, Vol. 4842,
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series, ed. E. Atad-
Ettedgui & S. D'Odorico, 139 -- 150
Garc´ıa Munoz, A., Zapatero Osorio, M. R., Bar-
rena, R., Montan´es-Rodr´ıguez, P., Mart´ın,
E. L., & Pall´e, E. 2012, ApJ, 755, 103
Husser, T.-O., Wende-von Berg, S., Dreizler, S.,
Homeier, D., Reiners, A., Barman, T., &
Hauschildt, P. H. 2013, A&A, 553, A6
Kaltenegger, L. & Traub, W. A. 2009, ApJ, 698,
Rogers, J. C., Apai, D., L´opez-Morales, M., Sing,
519
D. K., & Burrows, A. 2009, ApJ, 707, 1707
Kaltenegger, L., Traub, W. A., & Jucks, K. W.
2007, ApJ, 658, 598
Kasdin, J. 1995,
in Proceedings of the IEEE,
Vol. 83, Proceedings of the IEEE, 802 -- 827
Lowell, P. 1905, Lowell Obs. Bulletin., No. 17
Lutgens, F. & Tarbuck, E. 1995, The atmosphere:
an introduction to meteorology, The Atmo-
sphere: An Introduction to Meteorology (Pren-
tice Hall)
Origlia, L., Oliva, E., Maiolino, R., Gustafsson,
B., Piskunov, N., Kochucov, O., Vanzi, L., Min-
niti, D., Zoccali, M., Hatzes, A., & Guenther,
E. 2010, in Society of Photo-Optical Instrumen-
tation Engineers (SPIE) Conference Series, Vol.
7735, Society of Photo-Optical Instrumentation
Engineers (SPIE) Conference Series
Pasquini, L., Cristiani, S., Garc´ıa L´opez, R.,
Haehnelt, M., Mayor, M., Liske, J., Manescau,
A., Avila, G., Dekker, H., Iwert, O., Delabre,
B., Lo Curto, G., D'Odorico, V., Molaro, P.,
Viel, M., Vanzella, E., Bonifacio, P., di Mar-
cantonio, P., Santin, P., Comari, M., Cirami,
R., Coretti, I., Zerbi, F. M., Span`o, P., Riva,
M., Rebolo, R., Israelian, G., Herrero, A., Za-
patero Osorio, M. R., Tenegi, F., Carswell, B.,
Becker, G., Udry, S., Pepe, F., Lovis, C., Naef,
D., Dessauges, M., & M´egevand, D. 2010, in
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series, Vol. 7735,
Society of Photo-Optical Instrumentation En-
gineers (SPIE) Conference Series
Pavlov, A. A., Kasting, J. F., Brown, L. L., Rages,
K. A., & Freedman, R. 2000, JGR, 105, 11981
Pont, F., Zucker, S., & Queloz, D. 2006, MNRAS,
373, 231
Rodler, F., Kurster, M., & Barnes, J. R. 2013,
MNRAS, 432, 1980
Rodler, F., L´opez-Morales, M., & Ribas, I. 2012,
ApJl, 753, L25
Rodler, F., Kurster, M., & Henning, Th. 2008,
A&A, 485, 859
Rothman, L. S., Gordon, I. E., Barbe, A., Ben-
ner, D. C., Bernath, P. F., Birk, M., Boudon,
V., Brown, L. R., Campargue, A., Champion,
J.-P., Chance, K., Coudert, L. H., Dana, V.,
Devi, V. M., Fally, S., Flaud, J.-M., Gamache,
R. R., Goldman, A., Jacquemart, D., Kleiner,
I., Lacome, N., Lafferty, W. J., Mandin, J.-
Y., Massie, S. T., Mikhailenko, S. N., Miller,
C. E., Moazzen-Ahmadi, N., Naumenko, O. V.,
Nikitin, A. V., Orphal, J., Perevalov, V. I., Per-
rin, A., Predoi-Cross, A., Rinsland, C. P., Rot-
ger, M., Simeckov´a, M., Smith, M. A. H., Sung,
K., Tashkun, S. A., Tennyson, J., Toth, R. A.,
Vandaele, A. C., & Vander Auwera, J. 2009,
J. Quant. Spec. Radiat. Transf., 110, 533
Sagan, C., Thompson, W. R., Carlson, R., Gur-
nett, D., & Hord, C. 1993, Nature, 365, 715
Schindler, T. L. & Kasting, J. F. 2000, Icarus, 145,
262
Schneider, J. 1994, ApSs, 212, 321
Seifahrt, A., Kaufl, H. U., Zangl, G., Bean, J. L.,
Richter, M. J., & Siebenmorgen, R. 2010, A&A,
524, A11
Snellen, I. A. G., de Kok, R. J., le Poole, R., Brogi,
M., & Birkby, J. 2013, ApJ, 764, 182
Szentgyorgyi, A., Frebel, A., Furesz, G., Hertz,
E., Norton, T., Bean, J., Bergner, H., Crane,
J., Evans, J., Evans, I., Gauron, T., Jord´an,
A., Park, S., Uomoto, A., Barnes, S., Davis,
W., Eisenhower, M., Epps, H., Guzman, D.,
McCracken, K., Ordway, M., Plummer, D.,
Podgorski, W., & Weaver, D. 2012,
in So-
ciety of Photo-Optical Instrumentation Engi-
neers (SPIE) Conference Series, Vol. 8446, So-
ciety of Photo-Optical Instrumentation Engi-
neers (SPIE) Conference Series
Traub, W. A., Kaltenegger, M. C., & Jucks, K. W.
2008, SPIE, 7010, 51
Webb, J. K. & Wormleaton, I. 2001, Pasa, 18, 252
This 2-column preprint was prepared with the AAS LATEX
macros v5.2.
16
0.1
P
A
F
0.01
3σ
0.001
0
0.1 0.2 0.3 0.4 0.5 0.6 0.7
cut-off parameter
Fig. 5. -- False Alarm Probability (FAP) of the
Earth twin's transmission spectral signal detection
as a function of the cut-off parameter value. We
find that values of the cut-off parameter between ∼
0.29 and 0.43 give a FAP lower than 0.027, which
corresponds to a 3σ detection. The lowest value of
the FAP corresponds to a cut-off parameter value
of 0.4.
s
t
i
s
n
a
r
t
f
o
r
e
b
m
u
n
s
t
i
s
n
a
r
t
f
o
r
e
b
m
u
n
50
40
30
20
10
0
50
40
30
20
10
0
2
4
6
8
10
M3
M4
M5
M6
M7
M3
M4
M5
M6
M7
2
4
6
8
10
distance [pc]
Fig. 6. -- Number of transits required to mea-
sure the 760 nm O2 absorption band with 3σ con-
fidence versus distance of the planetary system
from Earth. As instrument, we assume a high-
resolution spectrograph mounted on the 39-m E-
ELT. The upper panel illustrates the results for a
G-CLEF-like instrumental layout which provides a
velocity span of 0.75 km s−1 pixel−1 and a spectral
resolution of 100, 000. The lower panel depicts the
results for a UVES-like spectrograph with velocity
span of 1.2 km s−1 pixel−1, an image slicer and a
spectral resolution of 110, 000. In these plots, we
assume no red noise. To account for different red
noise levels, the number of transits needs to be
multiplied by the values provided in Figure 8.
17
s
t
i
s
n
a
r
t
f
o
r
e
b
m
u
n
s
t
i
s
n
a
r
t
f
o
r
e
b
m
u
n
50
40
30
20
10
0
50
40
30
20
10
0
2
4
6
8
10
G-CLEF
M3
M4
M5
M6
M7
SIMPLE
M5
M6
M7
M8
M9
2
4
6
8
10
distance [pc]
Fig.
7. -- Top panel: Number of transits re-
quired to measure the 760 nm oxygen absorp-
tion band with G-CLEF with 3σ confidence ver-
sus distance of the planetary system from Earth.
Bottom panel: Same as above, but for measure-
ments of the 1268 nm oxygen absorption band
with the SIMPLE-spectrograph employing a spec-
tral resolving power of 100, 000. In these plots we
assume no red noise. To account for different red
noise levels, the number of transits needs to be
multiplied by the values provided in Figure 8.
8
6
4
2
r
o
t
c
a
f
e
m
i
t
0
0.0
0.2
0.4
0.6
0.8
1.0
contribution of red noise
Fig. 8. -- The effect of red noise in the number
of transits and the amount of time necessary to
detect O2 in the atmosphere of an Earth twin at a
3σ confidence level. For example, a red noise level
of 40% of the white noise level in the observed
spectra approximately doubles the observing time
required for a detection.
)
s
r
a
e
y
(
n
a
p
s
e
m
i
t
14
12
10
8
6
4
2
0
M8, 760 nm
M9, 760 nm
M8, 1270 nm
M9, 1270 nm
2
4
6
8
10
distance (pc)
Fig. 9. -- Time span of the observations versus
distance of the host star for observations of M8V
and M9V stars in the O2 A-band at 760 nm and
the O2 band at 1268 nm. The observations at 760
nm have a red noise level of 20%, while the obser-
vations at 1268 nm have a red noise level of 50% of
the white noise level. Unlike in the ideal case of no
red noise, the observations in the O2 band at 1268
nm are more efficient than the observations on the
O2 A-band only for M9V stars. These results are
based on SIMPLE and a G-CLEF-like instrument
mounted on the E-ELT.
18
s
r
a
t
s
f
o
r
e
b
m
u
n
30
25
20
15
10
5
0
0 - 10 pc
0 - 5 pc
1
2
3
4
5
6
7
8
9
spectral sub-type of M-dwarf
Fig. 10. -- Frequency of M-dwarfs in the solar
neighbourhood (distances d ≤ 10 pc) as a function
of spectral type. The most common sub-types are
M3 and M4 dwarfs. For distances up to 5 pc, the
most frequent sub-types are M3 and M5 dwarfs.
Fig.
11. -- Histogram of the distribution of
barycentric radial velocities for a sample of 176
M-dwarfs published by Chubak et al.
(2012) in
5 km s−1 bins.
. The bins of stars with veloci-
ties less than 15 and 10 km s−1, respectively, are
shown as dashed and square shaded areas in the
histogram. Assuming the Chubak et al. (2012)
sample is representative of all nearby M-dwarfs,
about 44.3% of the stars will have absolute radial
velocities smaller than 10 km s−1, and 56.8% will
have radial velocities smaller than 15 km s−1.
19
l
e
b
a
l
i
a
v
a
e
m
i
t
f
o
n
o
i
t
c
a
r
f
0.6
0.5
0.4
0.3
0.2
0.1
0
0
Vbary = 30 km/s
Vbary = 20 km/s
Vbary = 10 km/s
5
10
15
20
25
30
vr [km/s]
Fig. 12. -- Amount of time of optimum visibil-
ity of an object in units of the yearly observ-
ing time available at an ideal observing site as
a function of the absolute value of the systemic
radial velocity vr. We only consider the tar-
get to be visible when following criteria are ful-
filled: (a) air mass X ≤ 2, (b) Sun 15◦ below
horizon, and (c) a relative radial velocity of ±(15
to 30) km s−1. We note that the vast majority of
stars exhibit semi-amplitudes of the barycentric
velocity of Vbary > 20 km s−1.
|
1711.10699 | 1 | 1711 | 2017-11-29T06:35:13 | The Thermophysical Properties of the Bagnold Dunes, Mars: Ground-truthing Orbital Data | [
"astro-ph.EP"
] | In this work, we compare the thermophysical properties and particle sizes derived from the Mars Science Laboratory (MSL) rover's Ground Temperature Sensor (GTS) of the Bagnold dunes, specifically Namib dune, to those derived orbitally from Thermal Emission Imaging System (THEMIS), ultimately linking these measurements to ground-truth particle sizes determined from Mars Hand Lens Imager (MAHLI) images. In general, we find that all three datasets report consistent particle sizes for the Bagnold dunes (~110-350 microns, and are within measurement and model uncertainties), indicating that particle sizes of homogeneous materials determined from orbit are reliable. Furthermore, we examine the effects of two physical characteristics that could influence the modeled thermal inertia and particle sizes, including: 1) fine-scale (cm-m scale) ripples, and 2) thin layering of indurated/armored materials. To first order, we find small scale ripples and thin (approximately centimeter scale) layers do not significantly affect the determination of bulk thermal inertia from orbital thermal data determined from a single nighttime temperature. Modeling of a layer of coarse or indurated material reveals that a thin layer (< ~5 mm; similar to what was observed by the Curiosity rover) would not significantly change the observed thermal properties of the surface and would be dominated by the properties of the underlying material. Thermal inertia and grain sizes of relatively homogeneous materials derived from nighttime orbital data should be considered as reliable, as long as there are not significant sub-pixel anisothermality effects (e.g. lateral mixing of multiple thermophysically distinct materials). | astro-ph.EP | astro-ph | The Thermophysical Properties of the Bagnold Dunes, Mars: Ground-truthing Orbital
Data
Christopher S. Edwards1, Sylvain Piqueux2, Victoria E. Hamilton3, Robin L. Fergason4,
Ken E. Herkenhoff4, Ashwin R. Vasavada2, Kristen A. Bennett1, Leah Sacks5, Kevin
Lewis6, Michael D. Smith7
1Northern Arizona University, Department of Physics and Astronomy, NAU BOX 6010
Flagstaff, AZ 86011, USA, [email protected]
2Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA, USA
3Southwest Research Institute, Boulder, CO, USA
4U.S. Geological Survey, Astrogeology Science Center, Flagstaff, AZ
5Carelton College, Northfield, MN, USA
6Johns Hopkins University, Baltimore, MD, USA
7NASA Goddard Space Flight Center, Greenbelt, MD, USA
Key Points:
1) Thermally derived particle sizes of the Bagnold dunes from orbit and landed
assets are consistent with direct grain size measurements
2) Thermally derived particle sizes are not dramatically affected by surface ripples
or thin layers of induration/armoring
1
3) Sub-pixel mixing of sand with nearby materials likely resulted in overestimated
particle sizes in previous orbital measurements
Abstract: 244 words
Body Text + Figures: 8026
Body References: 78
Figures: 11
Tables: 1
2
1 Abstract
In this work, we compare the thermophysical properties and particle sizes derived
from the Mars Science Laboratory (MSL) rover's Ground Temperature Sensor (GTS) of
the Bagnold dunes, specifically Namib dune, to those derived orbitally from Thermal
Emission Imaging System (THEMIS), ultimately linking these measurements to ground-
truth particle sizes determined from Mars Hand Lens Imager (MAHLI) images. In general,
we find that all three datasets report consistent particle sizes for the Bagnold dunes (~110-
350 µm, and are within measurement and model uncertainties), indicating that particle sizes
of homogeneous materials determined from orbit are reliable. Furthermore, we examine
the effects of two physical characteristics that could influence the modeled thermal inertia
and particle sizes, including: 1) fine-scale (cm-m scale) ripples, and 2) thin layering of
indurated/armored materials. To first order, we find small scale ripples and thin
(approximately centimeter scale) layers do not significantly affect the determination of
bulk thermal inertia from orbital thermal data determined from a single nighttime
temperature. Modeling of a layer of coarse or indurated material reveals that a thin
layer (< ~5 mm; similar to what was observed by the Curiosity rover) would not
significantly change the observed thermal properties of the surface and would be
dominated by the properties of the underlying material. Thermal inertia and grain sizes
(e.g. lateral mixing of multiple thermophysically distinct materials).
of relatively homogeneous materials derived from nighttime orbital data should be
considered as reliable, as long as there are not significant sub-pixel anisothermality effects
3
2
Introduction
The surface of Mars has been characterized using thermal infrared observations
from orbit from the time of the Mariner and Viking orbiter missions (e.g. Neugebauer et
al., 1971). Viking Infrared Thermal Mapper (IRTM) (Chase et al., 1978) observations (e.g.
Christensen, 1982; Christensen, 1983, 1986; Kieffer et al., 1977; Kieffer et al., 1976)
permitted the physical characterization of broad regions of the Martian surface to be
constrained though little geologic context was available due to the large footprint of the
IRTM instrument (~25km). When the Thermal Emission Spectrometer (TES) (Christensen
et al., 2001) instrument onboard the Mars Global Surveyor arrived, new views of the
surface at much higher resolution (~3x6 km) permitted a host of additional thermophysical
investigations, which linked observations to local/regional scale geologic processes (e.g.
Bandfield & Edwards, 2008; Bandfield & Feldman, 2008; Mellon et al., 2000; Nowicki &
Christensen, 2007; Putzig & Mellon, 2007b; Putzig et al., 2005). However, with the arrival
of the Thermal Emission Imaging System (THEMIS) (Christensen et al., 2003) instrument
onboard the 2001 Mars Odyssey orbiter and its 100 m/pixel thermal infrared data
(Christensen et al., 2004), new investigations were made that linked outcrop-scale
compositional information to material physical properties (e.g. Bandfield, 2008; Bandfield
et al., 2013; Bandfield & Rogers, 2008; Christensen et al., 2003; Christensen et al., 2005;
Edwards et al., 2009; Edwards et al., 2014; Edwards et al., 2008; Edwards & Ehlmann,
2015; Hamilton & Christensen, 2005; Osterloo et al., 2008; Rogers et al., 2009; Rogers &
Bandfield, 2009; Rogers et al., 2005; Rogers & Fergason, 2011; Rogers & Nazarian, 2013;
Tornabene et al., 2008). These new investigations robustly characterized the geologic
4
histories and origins of surfaces on Mars.
The primary material property under consideration here is thermal inertia (TI),
defined as TI = (krc)1/2, where k is the thermal conductivity, r is the bulk density of the
material, and c is specific heat. TI is dominated by the bulk thermal conductivity on Mars
(Jakosky, 1986; Kieffer et al., 1973; Presley & Christensen, 1997a, 1997b, 1997c) and can
be related to an effective particle size for the upper several milli- to deci-meters of surface
material (e.g. Piqueux & Christensen, 2011; Presley & Christensen, 1997b, 1997c; Presley
& Craddock, 2006). The conversion from temperature to TI and ultimately to grain size
has been applied to both orbital and landed spacecraft data (Fergason et al., 2006a; Kieffer
et al., 1973). However, linking these effective derived particle sizes to in situ measurements
(i.e., ground-truthing) by correlating derived thermophysical properties of surface
materials with particle size distributions derived from imaging presents a significant
challenge, due to disparate instrument resolutions as well as sampling cadence and times.
This is especially true for those surfaces for which TI values which are not consistent with
loose particulate material (<~350 J m-2 K-1 s-1/2; e.g. Piqueux & Christensen, 2009a, 2009b)
or igneous bedrock (>1200 J m-2 K-1 s-1/2; e.g. Edwards et al., 2009). These intermediate
thermal inertia regimes (e.g. ~350-1200 J m-2 K-1 s-1/2) can be further complicated by the
non-homogeneity of the surfaces in question, which can introduce significant sub-pixel
anisothermality into a single temperature measurement (e.g. Putzig & Mellon, 2007a,
2007b). Additionally, the limited temporal sampling of most fixed-local-time polar orbiting
missions (e.g., those carrying TES and THEMIS) complicates the interpretation of
thermophysical signatures. The temporally limited measurements obtained by these
instruments do not capture the shape of the diurnal temperature response curve (e.g.
5
Fergason et al., 2006b), though a single nighttime temperature measurement can still
provide a strong lever arm to disambiguate the physical properties of the surface.
Importantly, over the lifetime of the Mars Odyssey mission the orbit migrated in local time
sampling from ~0300-0700 and 1500-1900, which provided additional diurnal coverage
over much of the martian surface.
The Mars Science Laboratory (MSL) Curiosity rover spent approximately 20 sols
(1222-1242) exploring and characterizing an active dune field in Gale Crater informally
named the Bagnold Dunes. Specifically, the rover was parked at a barchan dune informally
named "Namib" over these sols. The Bagnold dune field is located on the northwest flank
of Aeolis Mons (Fig. 1a), informally known as "Mt. Sharp", and provides an excellent
opportunity to link orbitally derived thermophysical properties to in situ observations, for
several reasons:
1) the active nature of this dune field limits the likelihood of cementation of (and
therefore vertical variation in) the particulate material;
2) aeolian dunes likely represent a relatively homogeneous material, reducing the
potential for material mixture induced anisothermality of the surface within a
given footprint;
3) the large areal extent of the dune field which is resolved by many THEMIS pixels
(100m/px); and
4) these dunes have particle sizes large enough (>35µm) to be characterized by
imaging instruments (e.g. ChemCam Remote Microscopic Imager (ChemCam
RMI) (Maurice et al., 2012; Wiens et al., 2012), and Mars Hand Lens Imager
(MAHLI) (Edgett et al., 2012)) onboard Curiosity.
6
In this work we link the orbital THEMIS thermophysical characteristics (and their
interpretation in terms of particle sizes) with ground observations using the thermal
infrared measurements of the Rover Environmental Monitoring Station (REMS) (Gómez-
Elvira et al., 2012) – Ground Temperature Sensor (GTS) (Hamilton et al., 2014; Sebastian
et al., 2010) to robustly determined particle sizes from Curiosity rover imagery. This inter-
dataset linkage can be used to better understand dunes and other landforms across the
surface of Mars and aid in interpretations of aeolian and geological processes.
3 Methods
3.1 THEMIS Observations
THEMIS has been in operation around Mars onboard the 2001 Mars Odyssey
spacecraft for ~15 years. Over this timeframe, it has imaged the majority of the planet
multiple times, both day and night (Edwards et al., 2011b), and has acquired many
observations of Gale Crater (Hamilton et al., 2014). THEMIS data were processed through
standard techniques that include the removal of time-dependent focal plane temperature
drift (Bandfield et al., 2004), temperature variations across the THEMIS calibration flag
(Edwards et al., 2011b), map projection and then the removal of band-dependent and band-
independent row- and column-correlated noise (Edwards et al., 2011b; Nowicki et al.,
2013). These processing techniques result in noise-corrected, calibrated radiance data.
THEMIS brightness temperatures were determined by matching the expected radiance
from a uniformly emitting (blackbody) source to the THEMIS-measured radiance at band
9 (Bandfield et al., 2004; Christensen et al., 2004; Edwards et al., 2011b). The brightness
temperature at band 9 (12.57µm) is taken to represent the surface temperature (Fergason et
al., 2006b) because the Martian atmosphere is relatively transparent in that band and the
7
highest signal to noise measurements are obtained with this band, especially for nighttime
data (Christensen et al., 2004; Edwards et al., 2009; Edwards et al., 2011b; Fergason et al.,
2006b). When deriving thermophysical properties, nighttime data (typically ~3-6 AM for
the Mars Odyssey orbit) are used as they minimize the effects of slopes and albedo,
permitting a more reliable TI determination from a single data point. In deriving THEMIS
thermal inertia values for this work (Figure 1B), the KRC thermal model (Kieffer, 2013)
was employed. We used this model with a host of input parameters such as the albedo,
atmospheric opacity, elevation (and corresponding surface pressure scaled based on the
seasonal evolution observed by the Viking Landers), emissivity, surface geometry (i.e.,
slope and azimuth), temperature dependent material properties (specific heat and thermal
conductivity) and modeled temperatures for a specified seasonal and local time range and
interval for the region of interest on the surface (see Kieffer, 2013 for a comprehensive
overview of the model capabilities). Previous work using the KRC thermal model with
THEMIS data determined these input parameters on a framelet (256-line segment parts of
the THEMIS image) basis (Fergason et al., 2006b) to facilitate the production of THEMIS
thermal inertia as a global dataset. For this study we have developed a method to derive
the thermal inertia using higher resolution input parameter datasets available at the MSL
landing site on a pixel by pixel basis (Figure 2B). This is similar to several other
investigations (e.g. Catling et al., 2006; Sefton-Nash et al., 2012), although those studies
did not use the publicly available KRC thermal model. Furthermore, we use the "full" KRC
model to develop a temperature to thermal inertia lookup table with 40 thermal inertia
values ranging from 10-2200 J m-2 K-1 s-1/2, which runs for two martian years prior to
outputting data in the third year to achieve seasonal stability. This is in contrast to previous
8
work that relied on the one-point mode which only runs a subset of seasonal cases
(Fergason et al., 2006b; Kieffer, 2013). Critically, our updated method, while
computationally inefficient, takes advantage of spatially registered data at the same scale
as the THEMIS IR data. Data that was input into our model included a Context Camera
(CTX) (Malin et al., 2007) Digital Terrain Model (DTM) used for the slope, azimuth, and
elevation (Figure 3B-D) of the individual pixel as well as THEMIS visible Lambert albedo
(V01494002; Edwards et al., 2011a) that is tied to TES Lambert albedo (Figure 3A). We
assumed a surface emissivity of 0.98 (relevant for Gale Crater (Bandfield et al., 2000)) and
we included latitude, longitude, and local time backplanes. The method described above
was applied to THEMIS image I17950012. This image was converted to band 9 brightness
temperature and resulted in 10s of millions of KRC model runs that generated temperature
to thermal inertia lookup tables with 40 thermal inertia nodes on a pixel by pixel basis
(~300K pixels). The look up table was then applied using linear interpolation to the
THEMIS temperatures. THEMIS image I17950012 has a local time of 04:12, a solar
longitude of 349.1, and a visible atmospheric dust opacity of 0.4 that was scaled to 6 mbar
(Smith et al., 2016) and was acquired in Mars year 27. This image was chosen due to its
ideal coverage of the site under investigation (e.g. the entire rover path present and future)
as well as its overall calibration quality (no data dropouts, limited time between the shutter
closing image and data acquisition, lack of saturated or undersaturated pixels, and no
evidence of enhanced line-line noise) (Bandfield et al., 2004; Christensen et al., 2004;
Edwards et al., 2011b).
Following the conversion to thermal inertia, THEMIS data were transformed to
particle size following the relationships established by (Presley & Christensen, 1997b,
9
1997c) and refined by (Piqueux & Christensen, 2011). This method uses a parameterization
of the model results of Piqueux and Christensen (2011) and accounts for an input surface
pressure (scaled for the elevation and season, ~800 Pa for Gale Crater), average upper layer
skin depth temperature (180K), thermal conductivity derived from thermal inertia, porosity
reasonable for particulate materials (40%; Denekamp & Tsur-Lavie, 1981), a solid density
of 2950 kg/m3 (consistent with basalt), and specific heat for basalt (595.5 J kg-1 K-1; Fujii
& Osako, 1973). We limit the scale of particle sizes to 1mm in Figure 2C because surfaces
with effective particle sizes greater than this can no longer be interpreted as strictly
particulate media (Piqueux & Christensen, 2009a, 2009b) and interpretations become non-
unique. Effective particle sizes larger than 1 mm are consistent with cemented or indurated
materials and without further context are non-unique. While further information can
certainly be drawn from the larger effective particle sizes, the focus of this paper is the
Bagnold sand dunes, and as such the need to consider coarser particle sizes was not
warranted.
3.2 GTS Observations
The GTS observations of interest span Sols 1222 through 1242 and were taken
while the Curiosity rover was parked with the GTS sensor pointed at the west-facing,
secondary slip face of Namib dune (Figure 4A). GTS data with the best calibration were
selected and grouped by local time bins (Figure 4B) to match the output of the KRC
thermal model used for analysis. Several measurement uncertainties exist when
considering the GTS data, including potential shadowing by the MSL mast within the
GTS observation footprint, and the influence of the radioisotope thermoelectric generator
(RTG). The effects of these uncertainties are likely difficult to quantify but may
10
contribute to up to a ~4K temperature error (Hamilton et al., 2014; Martínez et al., 2014).
A series of diurnal temperature curves were generated for a range of thermal inertia
values (varying from 100-350 J m-2 K-1 s-1/2 by 10 J m-2 K-1 s-1/2 increments), albedo
values (from 0.07-0.2 by 0.01 increments) with all the other appropriate inputs (e.g.
REMS-derived UV opacities scaled to 6.1 mbar resulting in 0.337 (Smith et al., 2016),
latitude, longitude, elevation, local time, assumed surface emissivity of 0.98, temperature
dependent material properties, etc.) using the KRC thermal model, for comparison with
measured temperatures from GTS data. Several metrics to assess the quality of the fit of
the GTS temperature data to the KRC model-derived output over the 21 sols were
computed including: 1) the RMS difference between the GTS and modeled data, 2) the
difference between the minimum and maximum diurnal temperature (∆T) of GTS and
modeled data, and 3) the difference of average diurnal temperatures between the GTS and
modeled data. We take three approaches to determine the albedo and thermal inertia in
order to compare methods. The first is a straightforward best fit by "eye", where the best
fit is qualitatively estimated. Second, the minimum RMS difference is used to
simultaneously solve for albedo and thermal inertia. And third, we use the minimum
difference of average diurnal temperatures to solve for albedo and then, using the derived
albedo, we find thermal inertia by using the minimum ∆T. The functional effect of the
thermal inertia is to change the amplitude of the diurnal curve, while albedo, which
controls the total amount of energy absorbed by the surface, results in a shift of the
average temperature of the diurnal curve (Figure 6). Each of these thermal inertia values
was then converted to particle size following the methods described above for THEMIS
data. This thermally derived albedo (0.08-0.13, Figure 5) is compared to orbitally derived
11
Lambert albedo from THEMIS visible imagery (0.15 is typical for the Bagnold dunes) to
ensure that a reasonable albedo is derived.
Several modeling approaches were taken in order to assess the sensitivity of the
derived thermal inertia to armoring/layering and variations in slope and azimuth in the
GTS footprint (Figure 5B & C). There are two slope peaks in the GTS footprint at ~7.5
and ~28 and azimuths range from 225-300 (Figure 5D), which potentially indicate that
a simple planar fit is not sufficient. The three approaches taken are as follows:
1) A best fit plane over the GTS footprint (Figures 4 & 5) is calculated and used
as the slope and azimuth input to the thermal model.
2) Each scene within the GTS footprint was discretized into ~5,000 individual
slope and azimuth combinations (facets) derived from Navcam stereo data.
The resulting diurnal temperatures were modeled for each thermal inertia and
albedo combination, using the additional inputs described above. The
sampling of these facets was calculated by evenly filling the GTS FOV in
angular space rather than in rectified spatial coordinates and intersecting these
vectors with the Navcam surface mesh, selecting the slope and azimuth
combination. This provides a reasonable approximation for the weighting of
the near- and far-field contributions to the measured radiance. After modeling
~5,000 unique diurnal curves, representing the range of temperatures
experienced by the surface for the given slope and azimuth distributions
(Figure 5), these temperatures are converted into radiance, averaged, and
converted back to a single brightness temperature per local time bin, and
allowed us to capture the effect of slopes at sub-pixel scales. This process
12
results in set of diurnal curves as a function of thermal inertia and albedo that
can then compared to the GTS data. While the resolution of the thermal inertia
and albedo lookup table used is somewhat coarse – increments of 10 J m-2 K-1
s-1/2 and 0.01 units of Lambert albedo – these incremental values result in
temperature variations that are likely within the uncertainty of the GTS
instrument (less than a few K).
3) Other scenarios were considered that match the fine-scale imagery of the dune
scuffs by the rover wheel. These images indicate a slightly indurated, thin
upper layer that has a lag of coarser particles. We model this scenario using a
layered thermal model in which the upper material thermal inertia, lower
material thermal inertia, and upper layer thickness can be varied
independently. Due to the non-unique results of this method, the
determination of the best fit was achieved by eye. We used a set of input
variables (e.g. layer thicknesses etc.) that included those derived from the
simple and facet models described above. The layered model is used in
conjunction with image derived particle sizes converted to thermal inertia
(185 J m-2 K-1 s-1/2 and 150 µm for the lower material and 225 J m-2 K-1 s-1/2
and 350 µm for the upper material) to simply assess the effects small scale
armoring/induration of the upper most layers has on the thermophysical
properties of dunes.
3.3 Particle sizes from Imagery
MAHLI data spanning the campaign of the Bagnold dunes (Table 1) were used to
determine the true particle sizes of the dune in question. Because each whole GTS
13
footprint was not analyzed with fine-scale rover imagery, given the relatively large GTS
footprint (meters scale), a set of representative MAHLI images were chosen for analysis.
These images cover the lee and stoss sides of a small-scale ripple. They also capture both
disturbed and undisturbed materials that may reveal armoring of finer grained materials
by a coarser grained surficial component (observed at MSL's "Rocknest" site and by
other missions; e.g. MER (Fergason et al., 2006a)), or surface crusts formed by weakly
cemented particulate materials (e.g. Piqueux & Christensen, 2009a, 2009b). Two
methods were employed to derive the particle size from MAHLI data (Figures 7 & 8)
which encompassed a range of standoff distances and thus resolutions:
1) Digital particle size analysis using a power spectral density function of
particle sizes determined using Morlet wavelets (Buscombe, 2013). This
method requires fully resolved particles and appropriate choice of input
parameters including sampling density, the relationship between the image
size (e.g. number of pixels) and the relative size of the particles (in pixels),
which ensures that there are sufficient points to analyze and that there are
sufficient octaves and notes to capture the size distribution of the particles,
respectively (Buscombe, 2013). The result of this algorithm is a particle size
distribution and the average particle size ± the standard deviation (Table 1;
Figure 8).
2) For a subset of MAHLI images (Figure 7), point counts were conducted by
hand. In this methodology, a grid of points was spaced evenly that resulted in
at least 100 measurement points was overlain on the image. The long and
short axes of the particle that lands under the given crosshair of a given grid
14
were measured and their average is reported as the particle's size. Crosshairs
that did not align with a particle were excluded from the point count survey
and crosshairs that align with an unresolvable particle are lumped into a
category that is defined by the minimum resolvable particle of the image (3x
the image resolution). The average and standard deviation of these particle
sizes are then determined from this distribution (Table 1; Figure 7).
4 Results
4.1 Particle sizes from Orbit
THEMIS data provide valuable medium resolution information about the
effective particle sizes of surfaces across vast regions of Mars at scales relevant for local
studies. These data leverage the thermophysical history and techniques of the Viking
IRTM (e.g. Christensen, 1983; Edgett & Christensen, 1991; Kieffer et al., 1977) and TES
(e.g. Jakosky et al., 2000; Mellon et al., 2000; Putzig & Mellon, 2007a; Putzig et al.,
2005) instruments. The particle size of aeolian dunes are of particular interest for several
reasons for this work. First, aeolian dunes provide the most uniform (from a particle size
perspective) surfaces to compare to landed observations in both visible imagery and
thermally (e.g. Fenton & Mellon, 2006; Fergason et al., 2006a; Fergason et al., 2006b)
and have a predictable grain size constrained by wind tunnel experiments under Martian
pressures (e.g. Greeley et al., 1980). Second, the Bagnold dunes represent the first time
that an active dune, rather than a mantled or inactive bedform, can be directly assessed
from orbit and ground. The Namib dune, in the southwest region of the main Bagnold
dune field where Curiosity stopped does not fill an entire THEMIS pixel, which would
also contain radiance from rock of the Murray and Stimson formations. This results in
15
mixing and anisothermality within a THEMIS pixel, ultimately reducing the utility of this
area for direct comparison with THEMIS. However, the remainder of the Bagnold dune
field, which fills many THEMIS pixels, is available for use and has similar characteristics
to the Namib dune. There is no indication that Namib dune is meaningfully different than
the remainder of the field from a particle size perspective and likely has a similar
sediment source and transport history. It is worth mentioning that slightly stronger
olivine-spectral features are observed in association with the barchan dunes, potentially
suggesting minor differences in sorting may be present (Laporte et al., submitted;
Viviano-Beck et al., 2014), though these minor spectral differences are difficult to
quantity from a physical properties perspective.
From analysis of THEMIS data (I17950012), the main Bagnold dune field
(observed by multiple THEMIS pixels) has a minimum apparent thermal inertia of 200 J
m-2 K-1 s-1/2, a maximum of 310 J m-2 K-1 s-1/2 and an average thermal inertia of 240 ± 20 J
m-2 K-1 s-1/2. This corresponds to a minimum particle size of 148 µm, a maximum of 968
µm and an average particle size of 251±84 µm. A single porosity for the entire image was
used (40%) and is likely relevant for the regolith dominated surfaces (e.g. Denekamp &
Tsur-Lavie, 1981; Piqueux & Christensen, 2011). The average particle size values were
calculated after the conversion of each pixel from thermal inertia to particle size and not
simply converting the average thermal inertia as the conversion is non-linear. These
thermal inertia and particle size values were calculated using the maximum extent of the
Bagnold dune field, while attempting to exclude areas subject to sub-pixel mixing of
basement bedrock. This result from the average Bagnold dune field contrasts with our
orbitally-derived particle sizes (using the same methods) of Namib dune which range
16
from 312 µm to 438 µm with an average of 396 ± 49 µm (Figure 2). The increased
particle sizes at Namib dune and the maximum values of the main Bagnold dunes field
likely represent the effect of subpixel mixing with the basement Murray and Stimson
formations which commonly have thermal inertia values > ~350 J m-2 K-1 s-1/2 translating
to effective particle sizes of >1 mm. This type of mixing is observed at scales (e.g. Figure
1B) finer than the 100m/px THEMIS images and has a dramatic effect on the average
derived particle sizes and thermal inertias.
4.2 Particle Sizes Derived from Landed Observations
4.2.1 GTS Derived Particle Sizes
As described above, several different scenarios were considered to model the GTS
data and derive particle sizes. The first being a simple model of a uniform thermal inertia
using the geometry of a best fit plane, and a second, complex facet-based model
incorporating small scale topography. The RMS best fit for both the plane and facet
models resulted in a thermal inertia of 200 J m-2 K-1 s-1/2. The second fitting metric
(minimum average and ∆T temperature differences) for the facet model resulted in a
thermal inertia of 170 J m-2 K-1 s-1/2, while the plane model resulted in a thermal inertia of
180 J m-2 K-1 s-1/2. By eye, we estimated the thermal inertia to be 190 J m-2 K-1 s-1/2. This
range of thermal inertias translate to particle sizes of ~106-216 µm and are functionally
equivalent to particle sizes derived from orbit (within the measurement and model
uncertainty). Neither the plane or facet models matched the time of peak temperature
recorded by the GTS (~30-45-minute mismatch); however, the facet model typically did
a better job at matching the time of peak temperature, with only a ~15-minute
discrepancy.
17
The albedo used in these models was not derived from another dataset, as is
common practice from orbit which reduces the number of free parameters, but was
instead permitted to vary. This additional free parameter resulted in a significant range of
best fit albedos (0.07-0.13), all of which are below the orbitally derived value of
~0.15±0.01 for the main Bagnold dune field.
4.2.2 Particle sizes from Imagery
Particle sizes determined from MAHLI imaging are treated as ground-truth,
especially those determined manually. There are significant target-to-target variations in
the derived particle sizes determined manually, for example the target Barby is
significantly coarser (~320-350 µm) than Otavi or the Gobabeb scoop sites (~150-200
µm). This is likely due to the sampling location with Barby, which was undisturbed near
the ripple crest of the stoss side of High Dune (nearby to Namib, Figure 1B), while the
other samples, including the Goabeb Scoop Site 2 discard (sieved) and Otavi were
acquired from the lee side of Namib dune proper. The grain size differences between the
lee and stoss sides of ripples/dunes are expected based on terrestrial experience and as
such likely samples a range of grain sizes expected of the active Bagnold dunes. The
values determined for the Gobabeb Scoop Site 2 discard and Otavi are most useful for
direct comparison to GTS derived particle sizes as the GTS footprint observed the area
near to the primary sampling region of Curiosity for the Bagnold dunes campaign (Figure
9).
The particle size values determined automatically via the wavelet method have
significantly larger standard deviations than those measured manually, often more than
±100 µm, which limits their utility to discriminate materials with approximately the same
18
particle sizes, in this case fine to medium sands. If these data were used solely without
manual point counts to verify the derived particle sizes in the cases measured here, the
automated method would likely overestimate the actual particle size (Table 1; Figure 8).
5 Discussion
5.1 Cross Dataset Comparisons
To first order, all the datasets examined in this work agree with one another, and
for the first time provide a quantitative linkage from measured particle sizes to orbital
data of an active dune field on Mars. Fine scale imagery resulted in particle sizes of
~150-350 µm (depending on the sampling location), the GTS data resulted in values
ranging from ~110 to 220 µm (considering the range of fitting and modeling methods)
and the orbital THEMIS data resulted in particle sizes of ~160 to 340 (from the average
and 1-s values).
However, there are important discrepancies. A much larger range of thermal
inertia values was determined from orbit for both the main dune field and Namib dune,
likely representing sub-pixel mixing of the higher thermal inertia basement Murray
mudstone and Stimson sandstone (Vasavada et al., 2017). This would artificially raise the
upper end of derived thermal inertia values which is not adequately captured in the
modeling from orbit, and thus bulk thermal inertia interpretations can represent a
significant problem for the fine-scale interpretation of future landing sites, as well as
geologic contexts. A large range in visible imagery-derived particle sizes was also
observed (Figure 8 & Table 1) that likely represents some amount of variability expected
to be present over much of the dune field. This variability did not seemingly have a large
effect on the derived thermal inertia values, which is discussed further in the next section.
19
The ~30 J m-2 K-1 s-1/2 differences in the GTS derived thermal inertia values, depending
on the type of fit conducted, are likely within or close to the uncertainty of both the
instrument and model (i.e., input parameters), and is also commonly expected from orbit.
However, the derivation of both thermal inertia and albedo from a single diurnal
curve can also provide non-unique results when considering measurement uncertainties.
While the average daily temperature is primarily controlled by the albedo (Figure 6B),
thermal inertia plays a secondary role (Figure 6A). The opposite is true for the diurnal
temperature range, where thermal inertia is the primary controlling factor, but albedo
plays a secondary role. As such the two variables are not completely independent and
cannot be perfectly separated given measurement uncertainties when solving for both
using a thermal model. The comparison of a thermally derived albedo to an independent
albedo determination from additional instrumentation acquired contemporaneously is
critical to reducing the free model parameters and providing better model fits to
complicated datasets.
5.2 Using Ground-truth to Interpret Orbital Data
GTS provides a critical link between ground-truth particle sizes and the globally
available, orbitally derived THEMIS thermal inertia covering ~100% of the Martian
surface at 100m/px from 60S to 60N. As is observed in other regions of Gale Crater
(Vasavada et al., 2017), the complex geological surfaces examined by GTS often result in
non-unique interpretations. Thermal inertia can be even more difficult to constrain from
orbit, with limited context, fewer constraints (typically full diurnal coverage is not
available), and larger footprints covering 104 m2 at best with current instrumentation (e.g.
THEMIS, TES, etc.). While sand dunes are likely the most thermophysically
20
straightforward features available on the surface of Mars, they still present significant
complications. Figures 4A and 5 illustrate the complex surfaces observed in Navcam
imagery where ripples are present at cm- to m-scales. Ripples such as these are observed
elsewhere on the dune field (e.g. Lapotre et al., 2016) and they complicate the
distribution of surface slopes and azimuths. Furthermore, as discussed earlier, the
variability in particle size across different sections (e.g., lee, stoss, trough, peak, etc.) of
the dune further complicates interpretations. However, the GTS scale footprint permits
the assessment of the effects of uncertainties on the thermal data, and thus, due to the
scale of the dunes observed here, permits a translation of the GTS findings to orbit for
use with interpreting THEMIS and TES data.
We found that the inclusion of small scale ripples (adding a slope and slope
azimuth distribution) in the thermal modeling of surfaces did not result in significantly
better fits to the measured surface temperatures (Figures 4 & 11). Some important
differences are apparent in Figure 11, namely related to the timing and fit of the post-
dawn data, which are better modeled when adding the distribution of sub-pixel slopes and
slope-azimuth facets. This has important implications for orbitally derived thermal
inertias (typically derived pre-dawn) where no differences are observed between simple
planar and facet model for fine-scale slopes of the magnitude observed at Namib dune.
This result indicates (and is borne out by our measurements) that the thermal inertia of
the dune derived from orbit is likely to be representative of the actual dune particle size
and does not require the consideration of fine (10s of cm scale) ripple features likely to be
prevalent on the surface of dunes.
21
The second consideration is the potential for armoring and/or induration of the
upper ~mm to ~cm of Martian dunes. This also does not likely represent a significant
impediment to orbitally derived thermal inertia values, especially for active Martian
dunes where relatively thin and only moderately coarser particle size armors have been
observed (e.g. Fergason et al., 2006a). While we were able to achieve slightly better fits
(e.g. better match the shape of the curve, and not just the nighttime or min/max
temperatures) by using a more complicated layered model, the effects, at least in the case
of Namib dune, are also very minor. By using the two measured particle sizes from Barby
(the armoring coarser layer, ~350µm) and Gobabeb (~150µm) as endmembers in this
layered model and varying the thickness of the layer (assuming an albedo of 0.11), we
find that layers ~a few mm (up to ~5 mm) thick behave almost identically to the
underlying material (Figure 10). A moderately coarse armor (such as that modeled here)
that is less than ~1 mm causes a maximum deviation of ~2K at local noon, which is
within typical instrument uncertainties reported for THEMIS (Christensen et al., 2004)
and GTS (Gómez-Elvira et al., 2012; Sebastian et al., 2010; Vasavada et al., 2017). These
coarser or indurated upper layers are more efficient at translating the heat to the
insulating base layer, which then controls the overall surface temperature (Figure 10). A
surface that has an armoring layer that is thicker than ~30 mm would be dominated by the
material properties of the upper layer and behave nearly the same a non-layered surface
composed of the same materials. However, armoring layers as thick as this have not been
observed on any bedforms by landed spacecraft, therefore these results suggest that the
thermal inertia derived from thermal observations likely represent the bulk, underlying
material and not the typically thin armoring material. These results are consistent with
22
observations of Saber bedform at the MER Spirit landing site, which was composed of a
finer-grained material covered by a coarser-grained armor. The modeling of Mini-TES
data found that a thin layer of coarser particles (e.g. 1-2 mm) did not significantly affect
the derived thermal inertial of aeolian bedforms. Similarly to the work presented here,
the surface temperature of the bedform was controlled primarily by the grain size of the
substrate and not the armoring lag (Fergason et al., 2006a). The fits of these significantly
more complicated models (both planar and faceted) do not affect pre-dawn temperatures
significantly (~1-2K) from a simple planar model, though they do in general fit the rise
and late afternoon/evening data better than a simple model (Figure 11). Similarly, this has
important considerations for the determination of particle sizes from orbit as the derived
thermal inertia, even with some armoring/induration (up to ~5 mm for the particle sizes
modeled here), is likely representative of the bulk particle size of the dune and wind
regime, rather than the thermal inertia of the lag, which is again borne out by the
THEMIS observations shown here.
5.3 Dunes and Sands on Mars
Thermophysical data from THEMIS and other instruments has enabled the
determination of the approximate bulk particle size of materials that comprise dunes on
Mars (Christensen, 1983; Edgett & Christensen, 1991; Fenton & Mellon, 2006; Fergason
et al., 2012; Fergason et al., 2006b). This has improved our understanding of the range of
materials mobile on Mars both now and in the recent past. It is instructive to compare the
particle size observed at Bagnold with those of other dunes observed globally, both from
orbital and surficial measurements, to understand how representative the Bagnold dune
field is to other regions on Mars. This comparison can then inform how the results
23
presented here can be applied more broadly to interpret the recent geologic and aeolian
history of not only Gale crater, but of the Martian surface.
Edgett and Christensen (1991) determined the average Martian dune particle size
of km-scale dune fields (Kaiser, Proctor, Rabe, and Moreux craters) using Viking
Infrared Thermal Mapper (IRTM) and Viking Orbiter images. These dunes have thermal
inertia values (converted to SI units using a scale factor of 41.86 from Viking-era units of
10-3 cal cm-2 K-1 s-1/2, and particle sizes updated from thermal inertia values following the
same methods from orbit used in this work and assuming a 6 mbar atmosphere) ranging
from 326 J K-1 m-2 s-1/2 to 356 J K-1 m-2 s-1/2, with an average thermal inertia of 343 J K-1
m-2 s-1/2. These thermal inertia values correspond to particle sizes greater than ~1000 µm
(coarse sand to very coarse sand) with our updated thermal inertia to particle size
conversion controlled by lab experiments and numerical models (e.g. Piqueux &
Christensen, 2009a, 2009b; Presley & Christensen, 1997b, 1997c) as compared to
previous conversion estimates of ~500µm (Edgett & Christensen, 1991). Edgett and
Christensen (1994) expanded the work of Edgett and Christensen (1991) by examining
IRTM thermal inertia data for dark crater floor units found in latitudes ±55°, correlating
thermal inertia values with surface morphology, and interpreting the sedimentary and
aeolian transport history of the materials. The average thermal inertia for dark intra-crater
features was 270 J K-1 m-2 s-1/2, and thermal inertia values ranged from 160 J K-1 m-2 s-1/2
to 440 J K-1 m-2 s-1/2. Fenton et al. (2003) derived particle size values from TES-derived
thermal inertia values for the Proctor crater dune. The average TES-derived thermal
inertia value is 277 ± 17 J m-2 K-1s-1/2, corresponding to an effective particle size of 600 ±
140 µm. Fenton and Mellon (2006) also calculated particle size values from TES-derived
24
thermal inertia values for the Proctor crater dune field. TES thermal inertia values range
between 260 J K-1 m-2 s-1/2 and 360 J K-1 m-2 s-1/2, corresponding to particles sizes ranging
between ~450 µm and >1 mm (medium to coarse sand). These estimations of dune
particle sizes are all comparable when considering the high uncertainties associated with
these calculations (Fergason et al., 2006b).
The thermal inertia and particle sizes of bedforms was also determined at both the
Mars Exploration Rover (MER) landing sites using Mini-TES data (Fergason et al.,
2006a). In many cases, Microscopic Imager (MI) data were also acquired of these
features, enabling the direct comparison of particle sizes derived from thermal inertia and
imagery in a manner similar to our work at Namib dune. At Meridiani Planum in the
bottom of Endurance crater, Mini-TES footprints from the Opportunity rover did not fall
cleanly on the stoss or lee side of the bedforms, and therefore all Mini-TES
measurements were averaged to determine the thermal inertia of the entire bedform. The
bedform at Endurance crater has a thermal inertia of 200 J K-1 m-2 s-1/2, which
corresponds to a particle size of ~150 µm (fine sand). The sand at this location was
deemed too hazardous to image directly with MI, therefore a different patch of sand
closer to the rover was instead imaged. Given the proximity, it was assumed that this
sand was likely similar to the bedform material observed by Mini-TES, as such making a
direct comparison between MI and Mini-TES reasonable. The particle size of the sand
was measured by hand in this location and was ~130 µm in diameter, similar to the values
that were thermally derived from Mini-TES (Fergason et al., 2006a) and likely within the
measurement/model error.
25
At the Gusev site, Mini-TES on the Spirit rover observed a bedform in Bonneville
crater that consisted of two distinct sections, which were each modeled separately. The
upper section was observed to climb the crater wall to the north and had a Pancam (Bell
et al., 2008) albedo of ~0.18. The second portion of the observed bedform was located on
the lower crater floor and had a Pancam albedo of ~0.23. The thermal inertia derived
from Mini-TES for the north crater wall bedform was 200 J K-1 m-2 s-1/2, corresponding to
a particle size of ~150 µm (fine sand) (Fergason et al., 2006a). The bedform near the
crater floor has a Mini-TES-derived thermal inertia of 160 J K-1 m-2 s-1/2, corresponding
to finer particle size of~60 µm (silt) (Fergason et al., 2006a). Rather than being composed
of different sands, Fergason et al. (2006a) suggested that the variable thermal inertia
between these two units may instead be caused by a thin layer (≤1 cm) of mantling
aeolian dust. This hypothesis is consistent with both the higher albedo and lower thermal
inertia values observed. Fergason et al. (2006a) further suggested that the darker
particles of the bedform section along the wall were consistent with enhanced bedform
activity and as such a less dusty surface. In our above work, we suggested that armoring
and/or induration of the upper ~cm of Martian dunes is not a significant factor in the
derivation of thermal inertia. The distinction with the observations of the bedforms in
Endurance crater is that they were mantled in dust, rather than a coarse-grained armor.
The finer-grained dust, given its smaller diurnal skin depth, has a much larger effect on
the derived thermal inertia (both from ground and orbital measurements) than a coarser-
grained armor, with a larger diurnal skin depth.
The thermal inertia derived for bedforms observed at the MER and MSL sites
(i.e., Bonneville and Endurance craters at MER and Bagnold dune field at MSL) are all
26
lower than thermal inertia values derived for intracrater bedforms from orbital IRTM-,
and TES-derived thermal inertia and particle size values (e.g. Aben, 2003; Edgett &
Blumberg, 1994; Edgett & Christensen, 1991; Fenton et al., 2003). However, the thermal
inertia values derived at Bonneville and Endurance craters (~200 J K-1 m-2 s-1/2) are the
same as the thermal inertias derived at the Bagnold dunes. The Bagnold dunes are likely
to be more active given their low albedos as compared to the Bonneville and Endurance
crater bedforms, where optically thick, but thermally thin dust may be present (e.g.
Fergason et al., 2006a). Given the modeling findings above and previous results we
conclude (e.g. Fergason et al., 2006a)it is likely that the Bonneville and Endurance dunes
share a similar bulk particle size as the Bagnold dunes as thin and patchy dust doesn't
dramatically affect the thermal inertia of these surfaces, and may have been transported
under similar regimes.
The large particle sizes determined in early orbital studies (e.g. coarse to very
coarse sands) are likely the direct function of the large measurement spot sizes used (km-
scale), rather than related to the bedforms themselves. As shown in this work, the small-
scale (cm- to m) features of bedforms do not dramatically affect the effective particle
sizes (Figure 4b & c), and rather it is the sub-pixel mixing with the substrate
(significantly enhanced as pixel sizes grows) that dominates the uncertainty in particle
size. While there is likely some variability among dune particle sizes on Mars, early
studies that rely on TES results almost certainly overestimate the particle size for all but
the largest dune fields.
27
6 Conclusions
The Bagnold dune field offers a unique opportunity to directly compare ground-
truth particles sizes derived from imagery and lander-scale thermophysical data to orbital
data at high enough spatial scales to resolve many Martian dune fields at geologically
relevant scales. The cross-comparison of particle size from the significantly different
methods presented in this work yield consistent results. This work shows that particle size
determination of homogeneous materials, despite the small-scale (i.e., cm- to m-scale)
features and thin armoring/induration of the upper surface (i.e., <1 cm), do not
dramatically affect the thermal inertia derivation and thus effective particle size
determination. Instead, we find that sub-pixel mixing of the substrate (e.g. bedrock)
dominates the particle size uncertainty, which can be reduced by acquiring data at finer
spatial scales. This effect likely drove the early interpretation of relatively coarse dune
particle sizes on Mars (coarse to very coarse sands) due to the reliance and availability of
km-scale TES and IRTM data. We propose that dune particle sizes on Mars are likely
more similar to those observed in the Bagnold dunes (fine-medium sands), rather than
those derive from early orbital studies.
28
7 Acknowledgements
The authors acknowledge the MSL team who operates the Curiosity rover on Mars,
as well as the REMS instrument team, that developed the Ground Temperature
Sensor. The authors thank Daniel Buscombe for assistance with the pyDGS software
package. Data used in this work was obtained from the NASA Planetary Data System.
Additional reduced data products are available from Christopher Edwards
([email protected]), NAU's Open Knowledge website
(http://openknowledge.nau.edu/id/eprint/5230), modeling and data analysis tools
are freely available (http://krc.mars.asu.edu and http://davinci.asu.edu). This
research was funded by the MSL Participating Scientist program and the 2001 Mars
Odyssey THEMIS instrument. Part of this work was performed at the Jet Propulsion
Laboratory, California Institute of Technology, under a contract with NASA.
29
8 Figure Legends
Figure 1. (A) Context Image mosaic with the rover traverse overlain, (B) HiRISE color
mosaic of the Bagnold dunes area. The orange star corresponds to the location where
activities associated with Namib dune from Sol 1222-1242 were undertaken.
30
Figure 2. (A) CTX basemap of lower Mount Sharp, including a portion of the rover
traverse path, the outline of the Bagnold dunes corresponding to our grain size results, and
an arrow that identifies Namib dune. (B) THEMIS thermal inertia derived from I17950012
of lower Mt. Sharp colorized and overlain on the CTX basemap. (C) THEMIS derived
particle size from I17950012 of lower Mt. Sharp colorized and overlain on the CTX
basemap. Particle sizes >1mm are not distinguishable and are colorized as white.
31
Figure 3. (A) THEMIS derived visible Lambert albedo scaled to 100m/px (Edwards et
al., 2011a), (B) CTX derived elevation scaled to 100m/px, (C) CTX derived slopes scaled
to 100m/px, and (D) CTX derived slope azimuth scaled to 100m/px used in the
derivation of the thermal inertia (Figure 2B) and ultimately particle size (Figure 2C).
32
Figure 4. (A) Navcam mosaic of Namib Dune. The approximate GTS footprint is highlighted in orange (B) GTS data
acquired over sols 1222-1242 along with several different model fits. (C) Difference of modeled and measured RMS
temperatures
for a range of albedo and thermal inertias. The locations of the model fits in (B) are shown.
33
Figure 5. Navcam derived (A) topography, (B) slopes, (C) slope azimuth of Namib
Dune. The approximate GTS footprint is highlighted in by the white outline. (D) A
normalized histogram showing the distribution of slopes and azimuths under the GTS
footprint. A best fit plane to the GTS footprint results in a slope of 17.7 and azimuth of
269.8.
34
Figure 6. Effects of (A) thermal inertia and (B) albedo over the ranges used in the
model fits. While the average diurnal temperature (dots) primarily depends on the
albedo of the surface, thermal inertia contributes as a second order effect complicating
the method of using average temperature to derive a "thermal" albedo.
35
Figure 7. MAHLI images at a variety of standoffs taken over the campaign at Namib
dune.
36
Figure 8. Bar graph of the average particle sizes determined from both by the Digital
Particle size analysis methods of Buscombe (2013)(dark gray) and by manual point
counts (light gray). Averages are shown with 1 sigma standard deviations. Manual
point counts have ~100-300 measurement location each.
37
Figure 9. Sampling location at Namib dune shown in the front left hazard camera.
Highlighted locations correspond to areas where MAHLI images in Table 1 and Figure
7 were acquired.
38
Figure 10. Thin layers of coarse or cemented material (225 J K-1 m-2 s-1/2, derived from
MAHLI particle sizes of Gobabeb) are modeled on top of a thermally thick layer of
lower thermal inertia (185 J m-2 K-1 s-1/2, derived from MAHLI particle sizes of Barby).
Layers of coarser material up to ~a few mm behave identically to the thermally thick
subsurface materials. Layers up to ~20mm are likely to be indistinguishable in GTS and
orbital data as the difference from a lower thermal inertia surface would be less than a
few K. Layers greater than a few cm (~30 mm) thick are dominated by the material
properties of the upper layer and behave nearly the same a non-layered surface
composed of the same materials. Surfaces with these characteristics would be
indistinguishable within instrument temperature uncertainties typical for GTS and
THEMIS. Full diurnal curves are shown in (A), while the difference from the 185 J K-1
m-2 s-1/2 material are shown in (B).
39
Figure 11. The best fit ∆T plane (180 J m-2 K-1 s-1/2; 0.13 albedo) and facet model (170 J
m-2 K-1 s-1/2; 0.12 albedo) data are compared to a layered plane and facet model with a
1 cm layer of 350 µm particles with an albedo of 0.12 on top of a 150 µm particle size
substrate. Full diurnal curves are shown in (A), while the difference between a simple
planar model and the other scenarios are shown in (B). In general, the most
complicated model (faceted and layered) provides the best fit of the data, matching the
early morning data, post-sunrise trends and the evening (~7 pm) data. No model used
in this work is able to accurately model either the peak temperature timing or the mid-
afternoon temperature drop, indicating there are some additional physical properties
not being adequately addressed.
40
Table 1.
41
9 References
Aben, L. K. (2003). Compositional and thermophysical analysis of Martian aeolian
dunes, M.S. thesis. Ariz. State Univ., Tempe, Dec., 117 pp.
Bandfield, J. L. (2008). High-silica deposits of an aqueous origin in western Hellas Basin,
Mars. Geophysical Research Letters, 35(12). doi:doi:10.1029/2008GL033807
Bandfield, J. L., & Edwards, C. S. (2008). Derivation of martian surface slope
characteristics from directional thermal infrared radiometry. Icarus, 193(1), 139-
157. doi:doi: 10.1016/j.icarus.2007.08.028
Bandfield, J. L., Edwards, C. S., Montgomery, D. R., & Brand, B. D. (2013). The dual
nature of the martian crust: Young lavas and old clastic materials. Icarus, 222(1),
188-199. doi:doi: 10.1016/j.icarus.2012.10.023
Bandfield, J. L., & Feldman, W. C. (2008). Martian high latitude permafrost depth and
surface cover thermal inertia distributions. J. Geophys. Res, 113, E08001.
doi:doi:10.1029/2007JE003007
Bandfield, J. L., Hamilton, V. E., & Christensen, P. R. (2000). A global view of Martian
volcanic compositions. Science, 287(5458), 1626-1630. doi:doi:
10.1126/science.287.5458.1626
Bandfield, J. L., & Rogers, A. D. (2008). Olivine dissolution by acidic fluids in Argyre
Planitia, Mars: Evidence for a widespread process? Geology, 36(7), 579-582.
doi:doi: 10.1130/G24724A.1
Bandfield, J. L., Rogers, A. D., Smith, M. D., & P.R.Christensen. (2004). Atmospheric
correction and surface spectral unit mapping using Thermal Emission Imaging
System data. J. Geophys. Res., 109, E10008. doi:doi: 10.1029/2004JE002289
Bell, J. F., Rice, M. S., Johnson, J. R., & Hare, T. M. (2008). Surface albedo observations
at Gusev Crater and Meridiani Planum, Mars. J. Geophys. Res., 113(E6). doi:doi:
10.1029/2007je002976
Buscombe, D. (2013). Transferable wavelet method for grain-size distribution from
Catling, D. C., Wood, S. E., Leovy, C., Montgomery, D. R., Greenberg, H. M., Glein, C.
images of sediment surfaces and thin sections, and other natural granular patterns.
Sedimentology, 60(7), 1709-1732. doi:doi: 10.1111/sed.12049
Mars. Icarus, 181(1), 26-51. doi:https://doi.org/10.1016/j.icarus.2005.10.020
R., & Moore, J. M. (2006). Light-toned layered deposits in Juventae Chasma,
Chase, S. C., Jr., Engel, J. L., Eyerly, H. W., Kieffer, H. H., Palluconi, F. D., &
Schofield, D. (1978). Viking infrared thermal mapper. Appl. Opt., 17(8), 1243-
1251.
Christensen, P. R. (1982). Martian dust mantling and surface composition: Interpretation
of thermophysical properties. J. Geophys. Res., 87(B12), 9985-9998. doi: doi:
10.1029/JB087iB12p09985
Christensen, P. R. (1983). Eolian intracrater deposits on Mars: Physical properties and
global distribution. Icarus, 56(3), 496-518. doi:doi: 10.1016/0019-
1035(83)90169-0
Christensen, P. R. (1986). The Spatial Distribution of Rocks on Mars. Icarus, 68(2), 217-
238. doi:doi:10.1016/0019-1035(86)90020-5
Christensen, P. R., Bandfield, J. L., Bell, J. F., III, Gorelick, N., Hamilton, V. E., Ivanov,
A., . . . Wyatt, W. (2003). Morphology and composition of the surface of Mars:
42
Mars Odyssey THEMIS results. Science, 300(2056), 2056-2061. doi:doi:
10.1126/science.1080885
Christensen, P. R., Bandfield, J. L., Hamilton, V. E., Ruff, S. W., Kieffer, H. H., Titus,
T., . . . Greenfield, M. (2001). Mars Global Surveyor Thermal Emission
Spectrometer experiment: description and surface science results. J. Geophys.
Res., 106(E10), 23,823-823,871. doi:doi:10.1029/2000JE001370
Christensen, P. R., Jakosky, B. M., Kieffer, H. H., Malin, M. C., McSween, H. Y., Jr.,
Nealson, K., . . . Ravine, M. (2004). The Thermal Emission Imaging System
(THEMIS) for the Mars 2001 Odyssey Mission. Space Science Reviews, 110(1),
85-130. doi:doi: 10.1023/B:SPAC.0000021008.16305.94
Christensen, P. R., McSween, H. Y., Bandfield, J. L., Ruff, S. W., Rogers, A. D.,
Hamilton, V. E., . . . Moersch, J. E. (2005). Evidence for magmatic evolution and
diversity on Mars from infrared observations. Nature, 436(7050), 504-509.
doi:doi: 10.1038/nature03639
Denekamp, S., & Tsur-Lavie, Y. (1981). Measurement of porosity in natural sand
deposits. Journal of Geotechnical and Geoenvironmental Engineering,
107(ASCE 16173).
Edgett, K., Yingst, R. A., Ravine, M., Caplinger, M., Maki, J., Ghaemi, F. T., . . . Goetz,
W. (2012). Curiosity's Mars Hand Lens Imager (MAHLI) Investigation. Space
Science Reviews, 170(1-4), 259-317. doi:doi: 10.1007/s11214-012-9910-4
Edgett, K. S., & Blumberg, D. G. (1994). Star and linear dunes on Mars. Icarus, 112,
448-464.
Edgett, K. S., & Christensen, P. R. (1991). The Particle-Size of Martian Aeolian Dunes.
Journal of Geophysical Research-Planets, 96(E5), 22765-22776. doi:doi:
10.1029/91je02412
Edgett, K. S., & Christensen, P. R. (1994). Mars aeolian sand: Regional variations among
dark-hued crater floor features. J. Geophys. Res., 99(E1), 1997-2018.
Edwards, C. S., Bandfield, J. L., Christensen, P. R., & Fergason, R. L. (2009). Global
distribution of bedrock exposures on Mars using THEMIS high-resolution
thermal inertia. J. Geophys. Res, 114, E11001. doi:doi: 10.1029/2009JE003363
Edwards, C. S., Bandfield, J. L., Christensen, P. R., & Rogers, A. D. (2014). The
formation of infilled craters on Mars: Evidence for widespread impact induced
decompression of the early martian mantle? Icarus, 228(0), 149-166. doi:doi:
10.1016/j.icarus.2013.10.005
Edwards, C. S., Christensen, P. R., & Hamilton, V. E. (2008). Evidence for extensive
olivine-rich basalt bedrock outcrops in Ganges and Eos chasmas, Mars. Journal of
Geophysical Research-Planets, 113(E11), E11003. doi:doi:
10.1029/2008je003091
Edwards, C. S., Christensen, P. R., & Hill, J. (2011a). Mosaicking of global planetary
image datasets: 2. Modeling of wind streak thicknesses observed in Thermal
Emission Imaging System (THEMIS) daytime and nighttime infrared data. J.
Geophys. Res., 116(E10), E10005. doi:doi: 10.1029/2011JE003857
Edwards, C. S., & Ehlmann, B. L. (2015). Carbon sequestration on Mars. Geology,
43(10), 863-866. doi:doi: 10.1130/g36983.1
Edwards, C. S., Nowicki, K. J., Christensen, P. R., Hill, J., Gorelick, N., & Murray, K.
(2011b). Mosaicking of global planetary image datasets: 1. Techniques and data
43
processing for Thermal Emission Imaging System (THEMIS) multi-spectral data.
J. Geophys. Res., 116(E10), E10008. doi: doi: 10.1029/2010JE003755
Fenton, L. K., Bandfield, J. L., & Ward, A. (2003). Aeolian processes in Proctor Crater
on Mars: Sedimentary history as analyzed from multiple data sets. Journal of
Geophysical Research: Planets, 108(E12).
Fenton, L. K., & Mellon, M. T. (2006). Thermal properties of sand from Thermal
Emission Spectrometer (TES) and Thermal Emission Imaging System (THEMIS):
Spatial variations within the Proctor Crater dune field on Mars. J. Geophys. Res.,
111, E06014. doi:doi:10.1029/2004JE002363
Fergason, R., Christensen, P., Golombek, M., & Parker, T. (2012). Surface properties of
the Mars Science Laboratory candidate landing sites: characterization from orbit
and predictions. Space Science Reviews, 1-35. doi:doi:10.1007/s11214-012-9891-
3
Fergason, R. L., Christensen, P. R., Bell III, J. F., Golombek, M. P., Herkenhoff, K. E., &
Kieffer, H. H. (2006a). Physical properties of the Mars Exploration Rover landing
sites as inferred from Mini-TES-derived thermal inertia. J. Geophys. Res., 111,
E02S02. doi: doi:10.1029/2005JE002583
Fergason, R. L., Christensen, P. R., & Kieffer, H. H. (2006b). High resolution thermal
inertia derived from THEMIS: Thermal model and applications. J. Geophys. Res.,
111, E12004. doi:doi: 10.1029/2006JE002735
Fujii, N., & Osako, M. (1973). Thermal diffusivity of lunar rocks under atmospheric and
vacuum conditions. Earth and Planetary Science Letters, 18(1), 65-71.
Gómez-Elvira, J., Armiens, C., Castañer, L., Domínguez, M., Genzer, M., Gómez, F., . . .
Martín-Torres, J. (2012). REMS: The Environmental Sensor Suite for the Mars
Science Laboratory Rover. Space Science Reviews, 170(1-4), 583-640. doi:doi:
10.1007/s11214-012-9921-1
Greeley, R., Leach, R., White, B., Iversen, J., & Pollack, J. (1980). Threshold windspeeds
for sand on Mars: Wind tunnel simulations. Geophysical Research Letters, 7(2),
121-124. doi:doI: 10.1029/GL007i002p00121
Hamilton, V. E., & Christensen, P. R. (2005). Evidence for extensive olivine-rich
bedrock in Nili Fossae, Mars. Geology, 33(6), 433-436. doi:doi:
10.1130/G21258.1
Hamilton, V. E., Vasavada, A. R., Sebastian, E., Juarez, M. D., Ramos, M., Armiens, C.,
. . . Zorzano, M. P. (2014). Observations and preliminary science results from the
first 100 sols of MSL Rover Environmental Monitoring Station ground
temperature sensor measurements at Gale Crater. Journal of Geophysical
Research-Planets, 119(4), 745-770. doi:doi: 10.1002/2013je004520
Jakosky, B. M. (1986). On the thermal properties of martian fines. Icarus, 66(1), 117-
124. doi:doi: 10.1016/0019-1035(1086)90011-90014
Jakosky, B. M., Mellon, M. T., Kieffer, H. H., Christensen, P. R., Varnes, E. S., & Lee,
S. W. (2000). The thermal inertia of Mars from the Mars Global Surveyor
Thermal Emission Spectrometer. J. Geophys. Res., 105, 9643-9652.
Kieffer, H. H. (2013). Thermal model for analysis of Mars infrared mapping. Journal of
Geophysical Research: Planets, 118(3), 451-470. doi:doi: 10.1029/2012je004164
Kieffer, H. H., Martin, T. Z., Peterfreund, A. R., Jakosky, B. M., Miner, E. D., &
Palluconi, F. D. (1977). Thermal and albedo mapping of Mars during the Viking
44
primary mission. J. Geophys. Res., 82, 4249-4291. doi:doi:
10.1029/JS082i028p04249
Kieffer, H. H., S.C. Chase, J., Miner, E., Munch, G., & Neugebauer, G. (1973).
Preliminary report on infrared radiometric measurements from Mariner 9
spacecraft. J. Geophys. Res., 78, 4291-4312. doi:doi: 10.1029/JB078i020p04291
Kieffer, H. H., S.C. Chase, J., Miner, E. D., Palluconi, F. D., Munch, G., Neugebauer, G.,
& Martin, T. Z. (1976). Infrared thermal mapping of the martian surface and
atmosphere: First results. Science, 193, 780-786.
Laporte, M. G. A., Ehlmann, B. L., Minson, S. E., Arvidson, R. E., Ayoub, F., Fraeman,
A. A., . . . Bridges, N. T. (submitted). Compositional Variationsin Sands of the
Bagnold Dunes at Gale Crater, Mars, from Visible-Shortwave Infrared
Spectroscopy and Comparisoin to Ground-Truth from the Curiosity Rover.
Journal of Geophysical Research-Planets, Bagnold Dunes Speical Issue.
Lapotre, M. G. A., Ewing, R. C., Lamb, M. P., Fischer, W. W., Grotzinger, J. P., Rubin,
D. M., . . . Yingst, R. A. (2016). Large wind ripples on Mars: A record of
atmospheric evolution. Science, 353(6294), 55-58. doi:doi:
10.1126/science.aaf3206
Malin, M. C., Bell, J. F., Cantor, B. A., Caplinger, M. A., Calvin, W. M., Clancy, R. T., .
. . Wolff, M. J. (2007). Context Camera Investigation on board the Mars
Reconnaissance Orbiter. J. Geophys. Res., 112, E05S04. doi:doi:
10.1029/2006JE002808.
Martínez, G. M., Rennó, N., Fischer, E., Borlina, C. S., Hallet, B., de la Torre Juárez, M.,
. . . Haberle, R. M. (2014). Surface energy budget and thermal inertia at Gale
Crater: Calculations from ground-based measurements. Journal of Geophysical
Research: Planets, 119(8), 1822-1838. doi:doi: 10.1002/2014je004618
Maurice, S., Wiens, R. C., Saccoccio, M., Barraclough, B., Gasnault, O., Forni, O., . . .
Vaniman, D. (2012). The ChemCam Instrument Suite on the Mars Science
Laboratory (MSL) Rover: Science Objectives and Mast Unit Description. Space
Science Reviews, 170(1-4), 95-166. doi:doi: 10.1007/S11214-012-9912-2
Mellon, M. T., Jakosky, B. M., Kieffer, H. H., & Christensen, P. R. (2000). High
resolution thermal inertia mapping from the Mars Global Surveyor Thermal
Emission Spectrometer. Icarus, 148(2), 437-455. doi:doi:10.1006/icar.2000.6503
Neugebauer, G., Munch, G., Kieffer, H., S.C. Chase, J., & Miner, E. (1971). Mariner
1969 infrared radiometer results: Temperatures and thermal properties of the
martian surface. Astron. J., 76, 719-728.
Nowicki, K., Edwards, C., & Christensen, P. (2013). Post-Projection Removeal of Row-
and Column-Correlated Noise in Line-Scanning Data: Application to THEMIS
Infrared Data. IEEE Transaactions - WHISPERS.
Nowicki, S. A., & Christensen, P. R. (2007). Rock abundance on Mars from the Thermal
Emission Spectrometer. J. Geophys. Res, 112, E05007.
doi:doi:10.1029/2006JE002798
Osterloo, M. M., Hamilton, V. E., Bandfield, J. L., Glotch, T. D., Baldridge, A. M.,
Christensen, P. R., . . . Anderson, F. S. (2008). Chloride-Bearing Materials in the
Southern Highlands of Mars. Science, 319(5870), 1651-1654. doi:doi:
10.1126/science.1150690
45
Piqueux, S., & Christensen, P. R. (2009a). A model of thermal conductivity for planetary
soils: 1. Theory for unconsolidated soils. J. Geophys. Res., 114, E09005.
doi:doi:10.1029/2008JE003308
Piqueux, S., & Christensen, P. R. (2009b). A model of thermal conductivity for planetary
soils: 2. Theory for cemented soils. J. Geophys. Res., 114(E09006).
doi:doi:10.1029/2008JE003309
Piqueux, S., & Christensen, P. R. (2011). Temperature-dependent thermal inertia of
homogeneous Martian regolith. J. Geophys. Res., 116, E07004. doi:doi:
10.1029/2011JE003805
Presley, M. A., & Christensen, P. R. (1997a). The effect of bulk density, particle shape,
and sorting on the thermal conductivity of particulate materials under martian
atmosphere pressures. J. Geophys. Res., 102(E4), 9221-9229. doi:doi:
10.1029/97JE00271
Presley, M. A., & Christensen, P. R. (1997b). Thermal conductivity measurements of
particulate materials, Part I: A review. J. Geophys. Res., 102(E3), 6535-6549.
doi:doi: 10.1029/96JE03302
Presley, M. A., & Christensen, P. R. (1997c). Thermal conductivity measurements of
particulate materials, Part II: Results. J. Geophys. Res., 102(E3), 6551-6566.
doi:doi: 10.1029/96JE03303
Presley, M. A., & Craddock, R. A. (2006). Thermal conductivity measurements of
particulate materials: 3. Natural samples and mixtures of particle sizes. J.
Geophys. Res, 111, E09013. doi:doi: 10.1029/2006JE002706
Putzig, N. E., & Mellon, M. T. (2007a). Apparent thermal inertia and the surface
heterogeneity of Mars. Icarus, 191(1), 68-94. doi:doi:
10.1016/j.icarus.2007.05.013
Putzig, N. E., & Mellon, M. T. (2007b). Thermal behavior of horizontally mixed surfaces
on Mars. Icarus, 191(1), 52-67. doi:doi: 10.1016/j.icarus.2007.03.022
Putzig, N. E., Mellon, M. T., Kretkea, K. A., & Arvidson, R. E. (2005). Global thermal
inertia and surface properties of Mars from the MGS mapping mission. Icarus,
173(2), 325-341. doi:doi: 10.1016/j.icarus.2004.08.017
Rogers, A. D., Aharonson, O., & Bandfield, J. L. (2009). Geologic context of bedrock
exposures in Mare Serpentis, Mars: Implications for crust and regolith evolution
in the cratered highlands. Icarus, 200(2), 446-462.
doi:doi:10.1016/j.icarus.2008.11.026
Rogers, A. D., & Bandfield, J. L. (2009). Mineralogical characterization of Mars Science
Laboratory candidate landing sites from THEMIS and TES data. Icarus, 203(2),
437-453. doi:doi: 10.1016/j.icarus.2009.04.020
Rogers, A. D., Christensen, P. R., & Bandfield, J. L. (2005). Compositional heterogeneity
of the ancient martian crust: Analysis of Ares Vallis bedrock the THEMIS and
TES data. J. Geophys. Res., 110, E05010. doi:doi:10.1029/2005JE002399
Rogers, A. D., & Fergason, R. L. (2011). Regional-scale stratigraphy of surface units in
Tyrrhena and Iapygia Terrae, Mars: Insights into highland crustal evolution and
alteration history. J. Geophys. Res., 116, E08005. doi:doi:10.1029/2010JE003772
Rogers, A. D., & Nazarian, A. H. (2013). Evidence for Noachian flood volcanism in
Noachis Terra, Mars, and the possible role of Hellas impact basin tectonics.
46
Journal of Geophysical Research: Planets, 118(5), 1094-1113. doi:doi:
10.1002/jgre.20083
Sebastian, E., Armiens, C., Gomez-Elvira, J., Zorzano, M. P., Martinez-Frias, J., Esteban,
B., & Ramos, M. (2010). The Rover Environmental Monitoring Station Ground
Temperature Sensor: a pyrometer for measuring ground temperature on Mars.
Sensors (Basel), 10(10), 9211-9231. doi:doi: 10.3390/s101009211
Sefton-Nash, E., Catling, D. C., Wood, S. E., Grindrod, P. M., & Teanby, N. A. (2012).
Topographic, spectral and thermal inertia analysis of interior layered deposits in
Iani Chaos, Mars. Icarus, 221(1), 20-42.
doi:https://doi.org/10.1016/j.icarus.2012.06.036
Smith, M. D., Zorzano, M.-P., Lemmon, M., Martín-Torres, J., & de Cal, T. M. (2016).
Aerosol optical depth as observed by the Mars Science Laboratory REMS UV
photodiodes. Icarus, 280, 234-248. doi:doi: 10.1016/j.icarus.2016.07.012
Tornabene, L. L., Moersch, J. E., McSween, H. Y., Hamilton, V. E., Piatek, J. L., &
Christensen, P. R. (2008). Surface and crater-exposed lithologic units of the Isidis
Basin as mapped by coanalysis of THEMIS and TES derived data products. J.
Geophys. Res., 113, E10001. doi: doi:10.1029/2007JE002988
Vasavada, A. R., Piqueux, S., Lewis, K. W., Lemmon, M. T., & Smith, M. D. (2017).
Thermophysical properties along Curiosity's traverse in Gale crater, Mars, derived
from the REMS ground temperature sensor. Icarus, 284, 372-386. doi:doi:
10.1016/j.icarus.2016.11.035
Viviano-Beck, C. E., Seelos, F. P., Murchie, S. L., Kahn, E. G., Seelos, K. D., Taylor, H.
W., . . . Morgan, M. F. (2014). Revised CRISM spectral parameters and summary
products based on the currently detected mineral diversity on Mars. Journal of
Geophysical Research: Planets, 119(6), 1403-1431. doi:10.1002/2014JE004627
Wiens, R. C., Maurice, S., Barraclough, B., Saccoccio, M., Barkley, W. C., Bell, J. F., . . .
Wong-Swanson, B. (2012). The ChemCam Instrument Suite on the Mars Science
Laboratory (MSL) Rover: Body Unit and Combined System Tests. Space Science
Reviews, 170(1-4), 167-227. doi:doi: 10.1007/S11214-012-9902-4
47
|
1612.06947 | 1 | 1612 | 2016-12-21T02:23:02 | Photometric observations of the mutual phenomena of the Galilean satellites at the Pulkovo observatory in 2014-2015 | [
"astro-ph.EP"
] | We present the results of photometric observations of the mutual phenomena in the system of Galilean satellites obtained during 2014-2015. The observations were performed using the 26-inch refractor, Normal Astrograph, ZA-320 telescope of the Pulkovo Observatory (084) and MTM-500 telescope at Pulkovo mountain station at Kislovodsk (C20). We made observations a total of 72 phenomena. We had derived 51 light curves of good and medium quality for 34 events. The RMS of determining the brightness is within a range from 0.02 to 0.19 mag, the average RMS is 0.06 mag. This work was supported by RFBR grant (project 15-02-03025). | astro-ph.EP | astro-ph |
Photometric observations of the
mutual phenomena of the Galilean
satellites at the Pulkovo observatory
in 2014-2015
M.Yu. Khovritchev1∗, I.S. Izmailov1,
E.A.Roshchina1, D.L.Gorshanov1, A.M. Kulikova1
1Pulkovo Observatory, Russian Academy of Sciences,
Pulkovskoe sh. 65, St. Petersburg, 196140 Russia
February 20, 2018
Abstract
We present the results of photometric observations of the mu-
tual phenomena in the system of Galilean satellites obtained during
2014-2015. The observations were performed using the 26-inch refrac-
tor, Normal Astrograph, ZA-320 telescope of the Pulkovo Observa-
tory (084) and MTM-500 telescope at Pulkovo mountain station at
Kislovodsk (C20). We made observations a total of 72 phenomena.
We had derived 51 light curves of good and medium quality for 34
events. The RMS of determining the brightness is within a range
from 0.02 to 0.19 mag, the average RMS is 0.06 mag. This work was
supported by RFBR grant (project 15-02-03025).
Introduction
Dynamical studies of the Galilean satellites require the series of accurate po-
sitional observations. An analysis of the light curves of satellites obtained
during the mutual phenomena allows us to get high-precise astrometric data
∗e-mail: [email protected]
1
Table 1: Telescopes, cameras and observers.
Telescope
Camera
Observers
Site and
code
Pulkovo
084
Pulkovo
084
Pulkovo
084
26-inch refractor,
D=650 mm,
F=10413 mm
Normal
Astrograph,
D=330 mm,
F=3500 mm
ZA-320,
D=320 mm,
F=3200 mm
FLIProline 9000,
FOV 12′
× 12′,
scale=0.24′′/pix
SBIG STL-11K,
× 23′,
FOV 35′
scale=0.53′′/pix
STL-
SBIG
16803,
FOV 39′
× 39′,
scale=0.6′′/pix
Number
of
light
curves
27
Izmailov I.S.
Khovrichev M.Yu.,
Kulikova A.M.
Gorshanov D.L.,
Petrova
S.N.,
Slesarenko V.Yu.,
K.N.,
Naumov
Ivanov
A.V.,
Sokov
E.N.,
Kupriyanov V.V.
Lyashenko A.Yu.,
Rusov S.A.
8
9
7
Mountain
Station
(Kislovodsk)
C20
MTM-500M,
D=500 mm,
F=6520 mm
STL-
SBIG
1001E,
FOV 21′
× 21′,
scale=1.2′′/pix
- differences of topocentric (in the case of mutual occultations) or heliocen-
tric (for mutual eclipses) coordinates of two satellites involved in the phe-
nomenon. Mutual phenomena in the system of Galilean satellites of Jupiter
are repeated with a period of six years and nine months. The aim of this work
is providing observational data for further analysis. Pulkovo Observatory was
taking part in the previous international campaigns of mutual phenomena ob-
servations in 1995, 1997, 2003 and 2009 (Emel'yanov et al (2011)). Pulkovo
Observatory had collect about of ten percents of all world observations dur-
ing the campaign of 2003 (Arlot et al
(2009)). Recent set of phenomena
was available for observations at Pulkovo (084) and at Mountain station
near Kislovodsk (C20) from October 2014 to August 2015. Jupiter's altitude
reached to 45 degrees in the meridian. Ephemeris of mutual phenomena were
provided by Celestial Mechanics Department of SAI MSU and presented at
the ephemeris server MULTI-SAT (Emel'yanov and Arlot (2008)).
2
Figure 1: Typical light curves of several phenomena.
3
1 Observations and data processing
Observations were carried out with three Pulkovo telescopes (26-inch re-
fractor, Normal Astrograph, astrograph ZA-320) and the MTM-500M tele-
scope of the Pulkovo Mountain Station near Kislovodsk (See Table 1). The
observations were performed according to the recommendations given in
Emel'yanov's paper (Emel'yanov (2008)). Observations began in advance
before the start of phenomenon and continued for up to the duration time
after it's end. The exposure time was varied from 0.1 s to 1 s, depending
on weather conditions and the magnitudes of satellites. No filter was used.
These CCD images were analyzed and integrated fluxes were calculated for
satellite images at the central moment of each frame. Dark current and flat
fields were taken into account by the standard way. As a result, we obtained
the ratio of satellite light flux and reference object or the difference between
magnitudes of the satellite and the reference object. We chose the satellite,
which did not involved in a phenomenon, or a background star as a refer-
ence object. To correct the background gradient due to planet's scattered
light we calculated the background intensities using a linear model accord-
ing to expression I(x, y) = Ax + By + C. The parameters A, B and C
were determined by the least squares method using intensity value of the
pixels (I(x, y)) within the ring-shaped area around the image of the satel-
lite. Whenever possible, the Jupiter was taken out of the field of view of the
camera to minimize errors in determining the background. The least radius
of aperture was selected.
2 The results and conclusions
In a total, we performed 73 observations of 37 mutual phenomena. Fifty-
series of CCD frames of suitable quality and light curves were obtained. Some
examples of light curves are shown in Fig. 1. The values of magnitude differ-
ences were approximated by a quadratic polynomials to estimate preliminary
values of photometric precision. The standard deviations of determining the
magnitudes are ranged from 0.02 to 0.19 mag. The obtained light curves are
available at the Pulkovo database http://puldb.ru/photometry/phemu2014-2015/.
Acknowledgments
This work was supported by Russian Foundation for Basic Research (project
no. 15-02-03025).
4
References
Emel'yanov, N. V. Special program of observations of Jovian and Saturnian
satellites for 2009. Solar System Research, 42, 5, 448 (2008).
Emel'yanov N. V., Arlot J.-E. The natural satellites ephemerides facility
MULTI-SAT. Astronomy and Astrophysics. 487, 759 (2008).
Arlot J.-E., Thuillot W., Ruatti C. et al, The PHEMU03 catalogue of ob-
servations of the mutual phenomena of the Galilean satellites of Jupiter.
Astronomy and Astrophysics, 493, 3, 1171 (2009).
Emelyanov N. V., Andreev M. V., Berezhnoi A. A. et al Astrometric results of
observations at Russian observatories of mutual occultations and eclipses
of Jupiter's Galilean satellites in 2009. Sol. Sys. Research, 45,3, 264 (2011).
5
|
1903.01206 | 1 | 1903 | 2019-03-04T12:28:25 | Flattened loose particles from numerical simulations compared to Rosetta collected particles | [
"astro-ph.EP"
] | Cometary dust particles are remnants of the primordial accretion of refractory material that occurred during the initial stages of the Solar System formation. Understanding their physical structure can help constrain their accretion process. We have developed a simple numerical simulation of aggregate impact flattening to interpret the properties of particles collected by COSIMA. The aspect ratios of flattened particles from both simulations and observations are compared to differentiate between initial families of aggregates characterized by different fractal dimensions $D_f$. This dimension can differentiate between certain growth modes. The diversity of aspect ratios measured by COSIMA is consistent with either two families of aggregates with different initial $D_f$ (a family of compact aggregates with fractal dimensions close to 2.5-3 and some fluffier aggregates with fractal dimensions around 2). Alternatively, the distribution of morphologies seen by COSIMA could originate from a single type of aggregation process, such as DLPA, but to explain the range of aspect ratios observed by COSIMA a large range of dust particle cohesive strength is necessary. Furthermore, variations in cohesive strength and velocity may play a role in the higher aspect ratio range detected (>0.3). Our work allows us to explain the particle morphologies observed by COSIMA and those generated by laboratory experiments in a consistent framework. Taking into account all observations from the three dust instruments on-board Rosetta, we favor an interpretation of our simulations based on two different families of dust particles with significantly distinct fractal dimensions ejected from the cometary nucleus. | astro-ph.EP | astro-ph | Astronomy & Astrophysics manuscript no. Lasue_agg_Rosetta_rev2_final
March 5, 2019
c(cid:13)ESO 2019
Flattened loose particles from numerical simulations compared to
Rosetta collected particles
J. Lasue1, I. Maroger1, R. Botet2, Ph. Garnier1, S. Merouane3, Th. Mannel4, 5, A.C. Levasseur-Regourd6, and M.S.
Bentley7
9
1
0
2
r
a
M
4
.
]
P
E
h
p
-
o
r
t
s
a
[
1
v
6
0
2
1
0
.
3
0
9
1
:
v
i
X
r
a
1 IRAP, Université de Toulouse, CNRS, CNES, UPS, Toulouse, France
e-mail: [email protected]
2 Université Paris-Saclay/Université Paris-Sud/CNRS, UMR 8502, LPS, Orsay, France
3 Max Planck Institute for Solar System Research, Göttingen, Germany
4 Space Research Institute of the Austrian Academy of Sciences, Graz, Austria
5 University of Graz, Graz, Austria
6 Sorbonne Université, CNRS, LATMOS, Paris, France
7 European Space Astronomy Centre, Madrid, Spain
Received September 15, 1996; accepted March 16, 1997
ABSTRACT
Context. Cometary dust particles are remnants of the primordial accretion of refractory material that occurred during the initial stages
of the Solar System formation. Understanding their physical structure can help constrain their accretion process.
Aims. The in situ study of dust particles collected at slow speeds by instruments on-board the Rosetta space mission, including
GIADA, MIDAS and COSIMA, can be used to infer the physical properties, size distribution, and typologies of the dust.
Methods. We have developed a simple numerical simulation of aggregate impact flattening to interpret the properties of particles
collected by COSIMA. The aspect ratios of flattened particles from both simulations and observations are compared to differentiate
between initial families of aggregates characterized by different fractal dimensions D f . This dimension can differentiate between
certain growth modes, namely the Diffusion Limited Cluster-cluster Aggregates (DLCA, D f ≈ 1.8), Diffusion Limited Particle-cluster
Aggregates (DLPA, D f ≈ 2.5), Reaction Limited Cluster-cluster Aggregates (RLCA, D f ≈ 2.1), and Reaction Limited Particle-cluster
Aggregates (RLPA, D f ≈ 3.0).
Results. The diversity of aspect ratios measured by COSIMA is consistent with either two families of aggregates with different initial
D f (a family of compact aggregates with fractal dimensions close to 2.5-3 and some fluffier aggregates with fractal dimensions around
2). Alternatively, the distribution of morphologies seen by COSIMA could originate from a single type of aggregation process, such
as DLPA, but to explain the range of aspect ratios observed by COSIMA a large range of dust particle cohesive strength is necessary.
Furthermore, variations in cohesive strength and velocity may play a role in the higher aspect ratio range detected (>0.3).
Conclusions. Our work allows us to explain the particle morphologies observed by COSIMA and those generated by laboratory
experiments in a consistent framework. Taking into account all observations from the three dust instruments on-board Rosetta, we
favor an interpretation of our simulations based on two different families of dust particles with significantly distinct fractal dimensions
ejected from the cometary nucleus.
Key words. comets: general -- comets: individual: 67P/Churyumov-Gerasimenko -- protoplanetary disks -- accretion: accretion disk
-- methods: numerical -- space vehicles: instruments
1. Introduction
1.1. Cometary dust particles
Comets are believed to preserve pristine dust grains and to pro-
vide information about their aggregation processes in the early
Solar System (e.g. Weidenschilling 1997; Blum 2000). Analyses
of data from the Giotto mission to comet 1P/Halley and of foil
impacts and aerogel tracks retrieved by the Stardust mission in
the coma of comet 81P/Wild 2 have indeed given clues to the
presence of low density dust particles built up of agglomerates,
possibly with different tensile strengths and porosities (e.g. Fulle
et al. 2000; Hörz et al. 2006; Burchell et al. 2008). The interpre-
tation of remote polarimetric observations of bright comets, such
as 1P/Halley and C/1995 O1 Hale-Bopp, has lead to similar con-
clusions (Levasseur-Regourd et al. 2008; Lasue et al. 2009). Ag-
gregation of solid particles in the early Solar System may there-
fore form a diversity of porosities represented by their fractal
dimension, D f (Dominik & Tielens 1997; Kempf et al. 1999;
Bertini et al. 2009). Understanding the structure of cometary dust
particles can give clues to these early Solar System processes
(Blum & Wurm 2008; Fulle & Blum 2017).
its
long
rendezvous with
1.2. The Rosetta mission
comet
During
26 month
67P/Churyumov-Gerasimenko (hereafter 67P)
in its 2015
apparition, the Rosetta spacecraft monitored the properties of
cometary dust particles released by the nucleus in the pre-
and post- perihelion phases, as well as during some outburst
events. Three instruments were specifically devoted to the study
of dust particles: i) COSIMA (the COmetary Secondary Ion
Mass Analyzer, Kissel et al. (2007)) collected dust particles
of 10 to 100 µm size on 1 cm2 targets, imaged them with a
microscope operating under grazing incidence illumination
Article number, page 1 of 9
A&A proofs: manuscript no. Lasue_agg_Rosetta_rev2_final
with a resolution of about 14 µm, and then analyzed them
through a mass spectrometer after indium ion beam ablation,
ii) MIDAS (the Micro-Imaging Dust Analysis System, Riedler
et al. (2007)) collected micron-sized dust particles on targets of
about 3.5 mm2, in order to obtain 3D images of their surfaces
down to tens of nanometers pixel resolution using atomic
force microscopy, and iii) GIADA (the Grain Impact Analyzer
and Dust Accumulator, Colangeli et al. (2007)) measured the
optical cross-section, speed, momentum and cumulative flux of
hundreds of sub-millimeter sized dust particles.
The COSIMA and MIDAS instruments collected dust par-
ticles at velocities in the 1 to 15 m s−1 range (Fulle et al. 2015),
that is to say at relative velocities much lower than the 6.1 km s−1
reached during the collection of 81P/Wild 2 samples. Their
chemical properties were thus mostly preserved, as well as part
of their physical structure. Some small particles, which could
be fragments of fragile individual particles, were nevertheless
noticed (e.g. Bentley et al. 2016; Merouane et al. 2016). Inter-
estingly enough, some particles appeared to be flattened, most
likely as a result of impact alteration (e.g. Langevin et al. 2016;
Mannel et al. 2016).
The Rosetta dust experiments provide complementary in-
sights into the properties of dust particles thanks to their dif-
ferent approaches (see, for a review, Levasseur-Regourd et al.
2018). As far as images are concerned, the total number of dust
particles detected is above 30,000 for COSIMA and above 1,000
for MIDAS (Levasseur-Regourd et al. 2018; Güttler et al. sub-
mitted).
More specifically, all images of dust particles indicate that
they consist of more or less porous agglomerates of smaller
grains (following the classification introduced in (Güttler et al.
submitted)). Their overall sizes,
identified by well-defined
boundaries, range from about 1 micrometer to tens of microm-
eters for MIDAS, and from tens of micrometers to several hun-
dreds of micrometers for COSIMA. The presence of aggregated
structures at distinct scales suggests a hierarchical aggregation
(Bentley et al. 2016). Indeed, the fractal dimension of a very
porous agglomerate detected by MIDAS was determined via a
density-correlation function (Mannel et al. 2016), to be equal to
1.7±0.1. Dust showers observed by GIADA were also explained
by the presence of fragile agglomerates with a fractal dimension
below 2, possibly disrupted through electrostatic fragmentation
induced by the spacecraft (Fulle et al. 2015, 2016). Consider-
ing fractal aggregation processes, the porosity of dust particles
in 67P/Churyumov-Gerasimenko can thus be estimated to be at
least equal to 90% for very porous ones, and about 75% for more
compact ones (e.g. Blum & Wurm 2008; Bertini et al. 2009).
The porosity of 67P/Churyumov-Gerasimenko's dust particles
has been estimated to be around 60% based on the density of
the nucleus and the composition measured by COSIMA (Fulle
et al. 2017). Analysis of the reflectance of porous dust particles
collected by COSIMA indicate that a high porosity (>50%) is
necessary to explain that the mean free path of photons in the
particle correspond to a significant fraction of the particle size
(Langevin et al. 2017).
Finally, it may be added that the properties of cometary dust
particles, as revealed by the Rosetta mission, are, as previously
suspected, remarkably comparable to CP-IDPs, i.e. Chondritic
Porous Interplanetary Dust Particles collected in the Earth's
stratosphere, and UCAMMs, i.e. UltraCarbonaceous Antarctica
MicroMeteorites collected in the snows of central regions of
Antarctica (e.g. Levasseur-Regourd et al. 2018).
The morphology, the structure and the composition of such
dust particles strongly suggest that, as well as cometary nuclei
Article number, page 2 of 9
Fig. 1. Diversity of crushed particle types detected by COSIMA (Nick:
compact particle, C; Alexandros: rubble pile, R; Estelle: shattered clus-
ter, S; and Johannes: glued cluster, G) (adapted from Langevin et al.
(2016)).
themselves, they formed in the solar nebula and the primordial
disk (e.g. Davidsson et al. 2016; Blum et al. 2017), and were
never processed within large objects.
1.3. Specificity of COSIMA results
COSIMA collected and analyzed cometary particles ejected by
67P/Churyumov-Gerasimenko on gold black covered targets
(Kissel et al. 2007). The dust particles ejected by the comet im-
pacted COSIMA targets at a speed <10 m s−1 according to GI-
ADA measurements (Rotundi et al. 2015) with a deceleration
<1 × 106 m s−2 according to Hornung et al. (2016). These val-
ues are enough to damage the initial structure of the dust parti-
cles during the collision, as visually assessed from the images
acquired by COSISCOPE after collection. With a resolution of
14 microns, the microscope enabled studies of particle typology
and flux (Langevin et al. 2016). The images show particles rang-
ing from a few tens to several hundreds of microns, the majority
of which appears to be built of micron-sized sub-components, as
confirmed by MIDAS (Bentley et al. 2016). Analysis of the par-
ticle morphologies identified four families of particles (Langevin
et al. 2016) which fall into two major classes, compact and clus-
tered. These families are :
1. Compact (type C) particles present well-defined bound-
aries without smaller satellite particles and with an apparent
height above the collecting plane of the same order of mag-
nitude as their horizontal (x and y) dimensions.
2. Shattered cluster (type S) particles are defined by clusters of
fragments for which no individual fragment makes up a ma-
jor fraction of the initial particle. These particles are inter-
preted as rearrangement of fragments within the impacting
particle without associated disruption.
3. Glued cluster (type G) particles have a well-defined shape
and a complex structure where sub-components appear to be
linked by a fine-grained matrix with a smooth texture.
4. Rubble piles (type R) particles comprise components much
smaller than their apparent size. Upon collision with the
J. Lasue et al.: Flattened loose particles compared to Rosetta
Fig. 2. Probability density of aspect ratio for each type of particles de-
tected by COSIMA (adapted from Langevin et al. (2016)).
plate, the sub-components rearranged themselves in a flat-
tened conical pile with many satellite components indicating
poor cohesion.
√
The different types of particles collected by COSIMA are illus-
trated in Fig. 1.
The grazing incidence illumination provided by COSIS-
COPE allows both the surface area of the collected particles and
their height (based on their projected shadow) to be determined
(see Fig. 1). The area is determined from the ratio of bright pixels
before and after exposure to the dust flux from the comet. An as-
pect ratio of the compacted particles can be obtained from height
area.
The aspect ratio density distribution for each detected particle
type is shown in Fig. 2. The compact particles, C, appear unbro-
ken and present the largest aspect ratios, with a first peak around
0.5 and another close to 1. The other particles present typical
aspect ratios of around 0.3 with the shattered clusters being the
flattest type of agglomerates. To understand the physical struc-
ture of cometary nuclei, it is important to infer, as far as possible,
properties of dust particles prior to their collection. COSIMA
analyses have shown a correlation between the flux of dust par-
ticles at various distances from the comet nucleus and their mor-
phology (Merouane et al. 2016). The fragmenting particles ap-
pear to have a mechanical strength of a few 1000 Pa (Hornung
et al. 2016) and their morphological diversity could result from
different collection speeds in the range from 1 m s−1 to 6 m s−1 as
investigated by laboratory simulations (Ellerbroek et al. 2017).
In this work, we investigate if different dust particle struc-
tures prior to their collection can also lead to the different mor-
phologies found by the Rosetta dust instruments. We present a
set of numerical simulations of fractal aggregates flattening on
impact with a plane surface, before presenting its results and
discussing their implications for the interpretation of the Rosetta
measurements.
2. Method
2.1. Fractal aggregates models
We expect the dust particles aggregating in the solar nebula to
present fractal structures. Fractal aggregates in the early Solar
System form a diversity of porosities that can be represented
by their fractal dimension, D f , based on their aggregation pro-
cesses (Wurm & Blum 1998). Aggregation simulations consider
Fig. 3. 3D representation of four aggregates representing the four differ-
ent aggregation processes considered in this work. DLCA (D f ≈ 1.8),
RLCA (D f ≈ 2.1), DLPA (D f ≈ 2.5) and RLPA (D f ≈ 3). A scale is
given in number of monomers, and a referential frame is indicated by
colored arrows (x=blue, y=red, z=green).
the collisions of spherical monomers which represent individual
grains aggregating to form dust particles (Güttler et al. submit-
ted). Four main aggregation processes, leading to significantly
different fractal dimensions, are considered: DLCA (D f ≈ 1.8),
RLCA (D f ≈ 2.1), DLPA (D f ≈ 2.5) and RLPA (D f ≈ 3). The
DL models are Diffusion Limited models, in which when one
monomer meets another one it sticks directly to it. The RL mod-
els are Reaction Limited models, in which molecular reactions
occur when two monomers encounter each other and result in
them sliding with respect to one another in order to maximize
the number of bonds, resulting in a more compact aggregate.
CA stands for Cluster-cluster Aggregation and PA for Particle-
cluster Aggregation : in PA particles, monomers are added to
the same main cluster which accretes all the mass and is rela-
tively compact, whereas in CA particles, monomers form sep-
arate clusters which then aggregate, thus resulting in a smaller
fractal dimension of the aggregate. The PA process occurs when
the number of monomers compared to the available volume is
high, increasing the chances of collision amongst small aggre-
gates.
Depending on the physical conditions of the primordial pro-
tosolar nebula, in terms of dust to gas ratio and dust compo-
sition, we can expect each of these kinds of aggregates to be
formed (Weidenschilling 1997; Kimura 2001). They have also
each been produced by computer simulations and laboratory
experiments simulating the initial stages of planetary accretion
(Meakin 1991; Blum & Wurm 2008).
2.2. Flattening simulation
In a first step, 3D off-lattice aggregates of a number N=10 000
identical spherical particles (called monomers) are generated ac-
cording to the 4 different aggregation processes described above.
The resulting fractal aggregates are characterized by different
Article number, page 3 of 9
A&A proofs: manuscript no. Lasue_agg_Rosetta_rev2_final
tions between the two spherical elements.
EVdW =
AHr
12d
≤ Ek =
2
3 πr3ρ(sin θ)2V2
(sin θ0)2 =
AH
8πr2ρdV2
(2)
(3)
where EVdW is the van der Waals energy, AH is the Hamaker
constant for the material considered, r is the radius of a sin-
gle monomer, d is the diameter of a monomer, Ek is the kinetic
energy of the monomer, ρ is the density of the monomer, θ is
the angle between the direction linking the two centers of the
monomers and the vertical direction, as illustrated in Fig. 4 and
V is the velocity of the aggregate with respect to the collecting
surface z = 0.9 The monomer diameter, d, is slightly larger (by
0.4 nm) than the steady state distance between the centers of two
touching monomers. Typical monomer diameters are considered
to be 20 nm or larger, making this difference negligible (< 2%).
We thus consider the distance between two touching monomer
centers to be equal to d. The Hamaker constant of two particles
interacting corresponds to a measure of the relative strength of
the particles material with respect to the attractive van der Waals
forces between them (Hamaker 1937).
We call θ0 the angle θ for which Eq (2) is an equality. sin θ0
is a threshold above which monomers may break their bonds
and bounce. Changing this parameter can either be viewed as
changing the cohesive strength between monomers or as chang-
ing the collection velocity, as Eq (3) shows. Thus, with the
Hamaker constant of dry minerals under vacuum conditions
AH ≈1 × 10−19 J (Israelachvili 2011):
-- sin θ0 ≈ 1 corresponds to very cohesive monomer bonds, low
collection speed or very small monomer size (value typically
obtained for V = 1 m s−1 and r =0.01 µm or for AH values
higher than 1 × 10−19 J)
-- sin θ0 ≈ 10−3 corresponds to all bonds being broken, rel-
atively higher collection speed, or larger monomer sizes
(value typically obtained for V =10 m s−1 and r =0.1 µm)
So, if a projected monomer meets another monomer, we can
). If ν ≤ 1, the monomer
compute a collision parameter (ν = sin θ
sin θ0
sticks to the one it bumps into. If ν > 1 , the incoming monomer
bounces according to a random direction based on the Lamber-
tian reflection rule (see Fig. 4) and sticks to the plane z = 0 or
previously stuck monomers if they are present.
To make the model more realistic with respect to potential
mass loss that may be incurred by the aggregates as they are flat-
tened and their bonds are broken, in further simulations a mass
loss probability Ploss is introduced. In this case, if a monomer
meets the condition ν > 1, then it will be removed from the
simulation with the mass loss probability Ploss which matches
the chance that some monomers do not stick to any others and
bounce back to free space during the collision.
An illustration of the effect of changing the value for sin(θ0)
is given in Fig. 5 where the morphology of flattened aggre-
gates is clearly dependent upon the initial structure of the ag-
gregates and the geometric parameters. Under conditions where
most bonds are broken (sin θ0 ≈ 10−3), the more compact aggre-
gates appear to generate a small pyramid of monomers with an
angle of repose. The more porous the aggregates, the flatter ap-
pears to be the projection. In the case where bonds are unbroken
(sin θ0 ≈ 1), similar structures appear, but some vertical chain-
like columns of monomers extending upwards are also present
Fig. 4. Representation of the collision geometry for 2 superposed
monomers. If θ < θ0, then the superposition remains stable. When
θ > θ0 the cohesive link between the monomers is broken and the upper
monomer will bounce randomly following some of the green arrows and
will attach itself to the z = 0 surface at a further point, thus fragmenting
the aggregate.
initial fractal dimensions according to the approximate relation-
ships: DLCA (D f ≈ 1.8), RLCA (D f ≈ 2.1), DLPA (D f ≈ 2.5)
and RLPA (D f ≈ 3). The values were calculated using the well
known self-similarity properties of fractals whereby the num-
ber, Nm, of monomers constituting the aggregate located within a
sphere of radius R follows Nm ∝ RD f , where R is smaller than the
gyration radius of the aggregate. The gyration radius of a frac-
tal aggregate is a measure of the extent of the aggregate, akin to
the standard deviation of the monomers' distance to the centre
of mass of the aggregate and can be calculated by
(ri − r j)2 =
1
N
(ri − rc)2
(1)
×(cid:88)
i, j
R2
g =
1
2N
×(cid:88)
i
where N is the number of monomers in the aggregate, ri and
r j are the spatial coordinates of the center of the monomers i
and j, and rc corresponds to the spatial coordinates of the center
of mass of the aggregate (Jullien & Botet 1987). A representa-
tion of each of the four aggregate types is given in Fig. 3. These
aggregates may correspond to different types of cometary parti-
cles as ejected from the surface of the nucleus by gas pressure.
For each aggregate type, 1000 different aggregation simulations
were performed to statistically analyze the results.
In a second step, simulating the particle collection and flat-
tening observed by COSIMA during the Rosetta mission, the
aggregates are projected monomer by monomer onto the plane
z = 0 as shown in Fig. 4. The monomers are selected iteratively
by increasing z values. If a monomer is projected directly onto
the plane z = 0 without encountering any other monomer, it
sticks directly to the collision plane. In the case where it encoun-
ters a monomer that is previously stuck under it, we consider that
a bond exists between the two monomers and that it will be bro-
ken if EVdW < Ek where EVdW is the van der Waals energy and
Ek is the kinetic energy of the incoming particle. This condition
can be written as in Eq (2) considering van der Waals interac-
Article number, page 4 of 9
J. Lasue et al.: Flattened loose particles compared to Rosetta
f
o
r
e
b
m
u
n
n
i
n
e
v
i
g
s
i
e
l
a
c
s
A
.
s
e
l
c
i
t
r
a
p
d
e
n
e
t
t
a
fl
e
h
t
f
o
y
g
o
l
o
h
p
r
o
m
e
h
t
n
o
n
o
i
s
e
h
o
c
d
n
o
b
d
n
a
)
f
D
(
.
)
n
e
e
r
g
=
z
,
d
e
r
=
y
,
e
u
l
b
=
x
(
s
w
o
r
r
a
d
e
r
o
l
o
c
y
b
d
e
t
a
c
i
d
n
i
s
i
e
m
a
r
f
l
a
i
t
n
e
r
e
f
e
r
a
d
n
a
,
s
r
e
m
o
n
o
m
y
g
o
l
o
h
p
r
o
m
e
l
c
i
t
r
a
p
l
a
i
t
i
n
i
f
o
t
c
e
ff
e
e
h
t
g
n
i
s
i
r
a
m
m
u
s
e
r
u
g
i
F
.
5
.
g
i
F
and increase the relative height of the flattened aggregate. These
columns of monomers appear due to the increased strength of the
bonds between the monomers, forming chain-like vertical struc-
tures that are not broken by the flattening geometry (as θ0 > θ
over the monomers' column). We therefore see that both param-
eters (D f and θ0) influence significantly the outcome of the sim-
ulated projection.
Fig. 6 shows the resulting projections in the case where the
monomers have a non-zero probability to bounce back to space
due to mass loss processes. In this case we only consider RLPA
aggregates with different Ploss values ranging from 0% to 50%.
As the mass loss probability gets larger, only a flat footprint of
the aggregate remains with a very low aspect ratio which can
represent the results of the low speed laboratory aggregate stick-
Article number, page 5 of 9
A&A proofs: manuscript no. Lasue_agg_Rosetta_rev2_final
Fig. 7. Illustration of the connected area calculated for a flattened ag-
gregate of type RLCA. The flattened particles seen from above is shown
on the left. The calculated connected areas are shown with gaps in the
middle and without gaps on the right. The parts of the aggregate that are
not connected to the largest connected aggregate are removed from the
processing.
Fig. 6. 3D view of a RLPA aggregate after projection with different
mass loss. A scale is given in number of monomers, and a referential
frame is indicated by colored arrows (x=blue, y=red, z=green) with an
approximate 30◦ viewing angle.
ing experiments of Ellerbroek et al. (2017) where most of the
initial aggregate mass was lost. Such mass loss processes may
also be at work during the COSIMA particle collection.
3. Results
3.1. Data Analysis
The aggregate flattening simulations were run to create 1000 ag-
gregates of each of the four types, using 10,000 monomers each,
for the four fractal dimensions considered, with the sin(θ0) pa-
rameter ranging from 1 to 10−6 and with a Ploss probability of
mass loss ranging from 0 to 0.5. This was done in order to ob-
tain good statistics for the aspect ratio of each numerically flat-
tened aggregate for comparison to the COSIMA measurements.
The height, H, of the flattened aggregates is the maximum value
of z among all the sticking monomers. To compute the area, A,
we considered only the monomers visible from above (looking
towards the −z direction) and, based on their position, we cal-
culated the contour of the projected connected set of monomers
(Lorensen & Cline 1987). We computed two different connected
areas: one with gaps and one without gaps as Fig. 7 shows. The
area with gaps is always somewhat smaller than the area with-
out gaps but is essentially linearly correlated with it. Therefore,
we calculated the results based on the connected area without
gaps. In this way, we can calculate a statistical distribution of
A, for particles of each kind similar to the
the aspect ratio, H/
procedure used with the COSIMA data and assess the effect of
the different parameters on the morphology of the flattened ag-
gregates.
√
3.2. The morphologies of flattened aggregates
Figure 8 represents the density distribution of aspect ratios cal-
culated for the 1000 flattened aggregates of each fractal type and
for 4 different values of sin(θ0). The upper figure is calculated
for a sin(θ0) = 1 corresponding to a simulation where no bond
Article number, page 6 of 9
between monomers is broken (illustrated on the right hand side
of the Figure 5). One can see that the distribution of aspect ratios
overlaps between about 0.5 (relatively flat aggregates) and 1.3.
The distribution also separates relatively well the different types
of aggregates with the more compact aggregates of type PA hav-
ing a median aspect ratio value of 1.18 (σ = 0.15) and the fluffier
aggregates of type CA having a median aspect ratio value of 0.73
(σ = 0.22). Therefore, to first order, the process appears to sep-
arate the aggregates with fractal dimensions above or below 2
into two groups. This is somewhat expected since more com-
pact aggregates will present more opportunities for solid verti-
cal structures of monomers to remain unbroken and to vertically
extend the projected aggregate. One can also notice that the as-
pect ratio distributions of the CA type aggregates present an ex-
tended right wing showing that some of those aggregates could
still have aspect ratios close to one, if their monomer bonds are
strong compared to the energy of impact.
As sin(θ0) decreases, the number of broken bonds increases
and the projected aggregates get flatter. The minimum aspect
ratio decreases and reaches 0.1 for values of sin(θ0) = 0.1 or
lower. The distribution of the most compact particles (RLPA
with D f ≈ 3) is now clearly separated from the distribution of
the other aggregates and remains around 1, indicating that the
surface dimensions covered by the flattened aggregate in x and
y are of the same order of magnitude as its vertical extent in z.
With respect to the distributions of the less compact aggregates,
we notice that the distributions for CA aggregates with fractal di-
mensions lower than about 2 become quickly undistinguishable.
Those flattened aggregates would therefore present essentially
the same aspect ratio distributions irrespective of their initial
morphology. The DLPA aggregates that have a fractal dimen-
sion around 2.5 are located in between those two extremes and
clearly separated from them at low values of sin(θ0). For exam-
ple, the standard deviation of the distributions for sin(θ0) = 0.25
range from 0.06 to 0.09. The DLPA distribution average aspect
ratio is approximately 0.3 for sin(θ0) = 0.1 or lower. At values
of sin(θ0) lower than 0.1, the density distributions stabilize to-
wards their final values. One can also notice a bimodal density
distribution for the flattened RLPA aggregates, corresponding to
whether vertical columns of monomers appear within the pyra-
mid somewhat extending its height. We expect the random size
distributions of monomers in real dust aggregates to limit the as-
pect ratio to the lower values of around 0.75-1.0. Some similar
linear chain-like structures were also detected in the analysis of
COSIMA particles, such as the 2CF Adeline particle (Hornung
et al. 2016).
The effect of the sin(θ0) parameter is further illustrated in
Figure 9 where aspect ratio distributions of DLPA aggregates are
J. Lasue et al.: Flattened loose particles compared to Rosetta
Fig. 9. Distribution of aspect ratio, H/
different sin(θ0) without mass loss.
√
A, for DLPA aggregates with
calculated for different values of sin(θ0) ranging from 0.001 to 1
and are superposed. As the sin(θ0) value decreases, the aspect ra-
tio decreases (due to the larger number of bonds breaking) from
approximately 1 to 0.25. One can also notice that the standard
deviation of the density distribution also decreases, indicating
that most aggregates of this type flatten in the same way. This is
related to the randomization of monomer deposition after bond
breaking which reduces the range of vertical extent possible after
flattening. Based on this figure, we can see that given a relatively
narrow range of collection velocities, since equation 3 indicates
that sin(θ0) is proportional to 1
V , a large range of bond cohesive
strengths in the aggregates would lead to a larger range of aspect
ratios for the same initial structure of the aggregate. This is espe-
cially true of DLPA as the aspect ratios for these particles range
from 0.25 to 1.2. Compact aggregate aspect ratios would range
between 0.8 and 1.3, while aggregates with fractal dimensions
around 2 and lower present aspect ratios ranging from 0.1 to 1.
If the cohesive strength of monomer bonds in the aggregate are
randomly distributed one can expect to detect more aggregates
with small flattened aspect ratios than large flattened aspect ra-
tios.
Finally, the effect of the mass loss coefficient is illustrated in
Figure 10 top where the probability density of aspect ratios for
RLPA aggregates with sin(θ0) = 1 is calculated for mass loss
parameters ranging from 0% to 50%. As expected, the mass loss
parameter reduces the aspect ratio of the flattened aggregates be-
cause of the loss of monomers. As compared to the variation in
aspect ratio distribution for varying sin(θ0), one can notice that
the end aspect ratio distribution remains relatively large (larger
than 0.5) which is due to the simultaneous loss of monomers in
all directions, so that the dimensions of the flattened aggregate
are reduced in all dimensions at more or less the same rate (in x,
y, and z). This parameter is also important in reducing the final
aspect ratio of the flattened aggregates.
In the case of DLPA aggregates and the more fluffy ones, the
initial aggregate is so porous that even moderate mass loss de-
stroys the structure during flattening. This leads rapidly to a very
flat final projected structure as illustrated in Figure 10 bottom.
It therefore appears that the initial fractal dimension of ag-
gregates strongly affects the morphology of flattened aggregates,
and that, depending on the effect of parameters such as the speed
of collection, strength of bonds between the monomers and mass
loss fraction, it may, or may not, be possible to distinguish the
initial structure of the particle from their flattened morphologies.
Article number, page 7 of 9
Fig. 8. Distribution of aspect ratio, H/
A, for 1000 aggregates of each
type (RLCA, DLCA, RLPA, DLPA) with sin(θ0) ranging from 1 to
0.001 without mass loss.
√
A&A proofs: manuscript no. Lasue_agg_Rosetta_rev2_final
2. the flattest kind of particles observed (shattered clusters with
an aspect ratio around 0.15) could be consistent with com-
paction of the smallest fractal dimension RLCA and DLCA
aggregates or with a very large mass loss during collection
(>50%).
3. the diversity of morphologies could also originate from a
single type of aggregation process (such as DLPA) but pre-
senting very different cohesive strengths amongst aggregates
(sin(θ0) ranging from at least 0.1 to 1). This distribution
would also present a peak around 0.3 as shown in Figure 9,
which would be consistent with the peak of the COSIMA
distribution around 0.3 as shown in Figure 2.
4. finally, a fourth process, described in Ellerbroek et al. (2017),
may be playing a role here as well. Experiments show that
incoming aggregates may sometimes fragment upon impact,
leaving some remains sticking to the target in a pyramidal
shape (mass transfer property between 0 and 0.8).
The diversity of aspect ratios observed appears consistent
with at least two families of aggregates with different D f , which
would also be consistent with the GIADA and MIDAS measure-
ments of two dust particles populations with very different frac-
tal dimensions, one being close to 3 and the other around 1.8
(Fulle & Blum 2017; Mannel et al. 2016)). Variations in both
the cohesive strength of the particles and the speed of collection
may play a role in the continuity of the higher aspect ratio range
(>0.3) detected by COSIMA. Alternatively, this could also mean
that the initial low fractal dimensions have been somewhat al-
tered by internal processes, such as compaction, or temperature
alteration, such as sintering, which may have happened during
the evolution of the cometary nucleus, especially on its surface.
Fig. 10. Distribution of aspect ratio, H/
A, for RLPA aggregates (top)
and DLPA aggregates (bottom) with sin(θ0) = 1 and different mass loss
probability coefficients.
√
4. Discussion
4.1. Comparison with COSIMA observations
The aspect ratio variation with initial D f
(aggregate type),
sin(θ0) and Ploss can be compared with the values observed by
COSISCOPE and presented in Fig. 2. On the one hand, only the
PA aggregate types have an aspect ratio large enough to explain
the presence of the compact particles in the COSIMA aspect ra-
tio distribution. This implies that a population of particles with
fractal dimension between 2.5 and 3 must be present in the dis-
tribution of particles ejected by 67P.
On the other hand, in order to explain the presence of mor-
phologies with aspect ratios as low as 0.1 to 0.3, where the dis-
tributions of COSIMA particles of type G, R and S peak, other
types of particles or processes need to be invoked. From our sim-
ulations, even with a mass loss as large as 50%, RLPA aggregates
alone cannot explain the range of aspect ratio observed. How-
ever, the DLPA type particles could reach aspect ratio values as
low as 0.2 either with different cohesive strengths and/or veloc-
ities (sin(θ0)) or with mass losses up to 50%. Finally, a fractal
dimension lower than 2 would also lead to very low final as-
pect ratios even when considering particles with higher cohesive
strengths. The large range of distribution observed by COSIMA
could therefore be explained by:
1. two different initial groups of particles with low and high
fractal dimensions (such as RLPA for the compact particles
and DLPA for the shattered clusters).
Article number, page 8 of 9
4.2. Comparison with collision experiments
In the work of Ellerbroek et al. (2017), laboratory simulations of
impacts of aggregates simulating the particle collection proce-
dure of Rosetta were presented. The aggregates were formed by
aggregation of irregular polydisperse SiO2 particles with density
around 2.6 kg m−3 and a size range of 0.1 to 10 µm. The final
aggregates have porosities around 65% ± 5% and low compres-
sive strength between 1 × 104 Pa and 1 × 106 Pa. The aggregates
were then accelerated by electrostatic forces towards a collecting
plane where the collision was filmed and the resulting flattened
footprint imaged and analyzed. The velocity of impact ranges
from about 1 m s−1 to 6 m s−1.
The footprints obtained represent the diversity of morpholo-
gies that were acquired by the COSIMA instrument. At very low
velocities of around 1 m s−1, the aggregates either stick directly
to the surface, similar to the compact COSIMA particle type,
or they may bounce from the surface, leaving a very flat foot-
print with mostly unconnected fragments, possibly morphologi-
cally similar to the shattered cluster COSIMA type of particles.
As velocities are increased from 2 m s−1 to 6 m s−1, the particles
mostly stick to the surface and fragmentation occurs, leading to
footprints morphologically similar to COSIMA rubble piles or
glued clusters.
In this laboratory work, all morphologies were generated us-
ing only a change in the impact velocity and impactor size, and
similarities could be seen between the footprints of the parti-
cles that were obtained on the collecting surface and the mor-
phologies measured by COSIMA. The simulations presented in
our work allow us to generate similar conditions of flattening by
varying the velocity and the particles sizes. However, in our sim-
ulations, we can also modify the initial impacting particle mor-
phology and study its effect on the flattening of the aggregates.
J. Lasue et al.: Flattened loose particles compared to Rosetta
This allows us to explore an extended set of parameters com-
pared with the laboratory experiments, and we have shown that
it is also possible to generate the measured footprint morphology
by considering different initial fractal dimensions of the impact-
ing particles, as discussed above. It would be of interest to study
in the laboratory how very porous particles behave when sub-
jected to the type of collection that happened during the Rosetta
mission to confirm our analysis.
4.3. Possible analysis of MIDAS data
A planned future study aims to investigate whether these results
are also valid for MIDAS particles. The aspect ratios of dust par-
ticles collected by MIDAS should be calculated and their distri-
bution reviewed. It will be of great interest if the distribution falls
in different groups, and if they match those found in the simu-
lation and with COSIMA particles. As MIDAS particles are one
order of magnitude smaller than those of COSIMA, this will al-
low us to understand how the initial structures of dust particles of
comet 67P might look and if they remain similar over the 1 µm
to 100 µm size range.
5. Conclusions
In this work, we have shown that simple numerical simulations
of aggregate flattening can be used to infer the initial proper-
ties of particles collected by COSIMA on-board Rosetta. The
diversity of aspect ratios measured in COSIMA images appears
consistent with several hypotheses on the initial properties of the
collected particles.
1. It could be explained by at least two families of aggregates
with different fractal dimensions D f . A mixture of some
compact particles with fractal dimensions close to 2.5-3 to-
gether with some fluffier ones with fractal dimensions <2
would also be consistent with the observations and the mea-
surements made by GIADA and MIDAS (Fulle & Blum
2017; Mannel et al. 2016).
2. Alternatively,
the distribution of morphologies seen by
COSIMA could originate from a single type of aggregation
process, such as DLPA (D f ≈ 2.5) but presenting a large
range of cohesive strengths or collection velocities. This dis-
tribution would be consistent with a maximum at an as-
pect ratio around 0.3 as observed on the COSIMA typol-
ogy (Langevin et al. 2016).
Furthermore, variations in cohesive strength and velocity may
play a role in the higher aspect ratio range detected by COSIMA
(>0.3). Our work allows us to explain the particle morphologies
observed by COSIMA and those generated by the laboratory ex-
periments of Ellerbroek et al. (2017) in a consistent framework.
Taken together with the observations made by GIADA and MI-
DAS on Rosetta, our simulations seem to favor an interpreta-
tion based on two different families of dust particles with sig-
nificantly distinct fractal dimensions ejected from the cometary
nucleus.
Acknowledgements. The authors acknowledge two anonymous referees for their
positive evaluation and constructive comments. The authors acknowledge sup-
port from Centre National d'Etudes Spatiales (CNES) in the realization of in-
struments devoted to space exploration of comets and in their scientific analysis.
T.M. acknowledges funding by the Austrian Science Fund FWF P 28100-N36.
References
Bentley, M. S., Schmied, R., Mannel, T., et al. 2016, Nature, 537, 73
625
Bertini, I., Gutierrez, P. J., & Sabolo, W. 2009, Astronomy & Astrophysics, 504,
Blum, J. 2000, Space Science Reviews, 92, 265
Blum, J., Gundlach, B., Krause, M., et al. 2017, Monthly Notices of the Royal
Astronomical Society, 469, S755
Blum, J. & Wurm, G. 2008, Annu. Rev. Astron. Astrophys., 46, 21
Burchell, M. J., Fairey, S. A., Wozniakiewicz, P., et al. 2008, Meteoritics & Plan-
Colangeli, L., Lopez-Moreno, J. J., Palumbo, P., et al. 2007, Space Science Re-
Davidsson, B. J. R., Sierks, H., Guettler, C., et al. 2016, Astronomy & Astro-
etary Science, 43, 23
views, 128, 803
physics, 592, A63
Dominik, C. & Tielens, A. 1997, The Astrophysical Journal, 480, 647
Ellerbroek, L. E., Gundlach, B., Landeck, A., et al. 2017, Monthly Notices of the
Royal Astronomical Society, 469, S204
Fulle, M. & Blum, J. 2017, Monthly Notices of the Royal Astronomical Society,
Fulle, M., Della Corte, V., Rotundi, A., et al. 2017, Monthly Notices of the Royal
Fulle, M., Della Corte, V., Rotundi, A., et al. 2015, The Astrophysical Journal
469, S39
Astronomical Society, 469, S45
Letters, 802, L12
Fulle, M., Levasseur-Regourd, A. C., McBride, N., & Hadamcik, E. 2000, The
Astronomical Journal, 119, 1968
Fulle, M., Marzari, F., Della Corte, V., et al. 2016, The Astrophysical Journal,
Güttler, C., Mannel, T., Rotundi, A., et al. submitted, Astronomy & Astrophysics
Hamaker, H. C. 1937, physica, 4, 1058
Hornung, K., Merouane, S., Hilchenbach, M., et al. 2016, Planetary and Space
821, 19
Science, 133, 63
Hörz, F., Bastien, R., Borg, J., et al. 2006, science, 314, 1716
Israelachvili, J. N. 2011, Intermolecular and surface forces (Academic press)
Jullien, R. & Botet, R. 1987, Ann. Telecomm., 41, 343
Kempf, S., Pfalzner, S., & Henning, T. K. 1999, Icarus, 141, 388
Kimura, H. 2001, Journal of Quantitative Spectroscopy and Radiative Transfer,
Kissel, J., Altwegg, K., Clark, B. C., et al. 2007, Space Science Reviews, 128,
70, 581
823
Langevin, Y., Hilchenbach, M., Ligier, N., et al. 2016, Icarus, 271, 76
Langevin, Y., Hilchenbach, M., Vincendon, M., et al. 2017, Monthly Notices of
the Royal Astronomical Society, 469, S535
Lasue, J., Levasseur-Regourd, A. C., Hadamcik, E., & Alcouffe, G. 2009, Icarus,
199, 129
reviews, 214, 64
Science, 56, 1719
Levasseur-Regourd, A.-C., Agarwal, J., Cottin, H., et al. 2018, Space science
Levasseur-Regourd, A.-C., Zolensky, M., & Lasue, J. 2008, Planetary and Space
Lorensen, W. E. & Cline, H. E. 1987, in ACM siggraph computer graphics,
Vol. 21 (ACM), 163 -- 169
Mannel, T., Bentley, M. S., Schmied, R., et al. 2016, Monthly Notices of the
Royal Astronomical Society, stw2898
Meakin, P. 1991, Reviews of Geophysics, 29, 317
Merouane, S., Zaprudin, B., Stenzel, O., et al. 2016, Astronomy & Astrophysics,
Riedler, W., Torkar, K., Jeszenszky, H., et al. 2007, Space Science Reviews, 128,
596, A87
869
Rotundi, A., Sierks, H., Della Corte, V., et al. 2015, Science, 347, aaa3905
Weidenschilling, S. J. 1997, Icarus, 127, 290
Wurm, G. & Blum, J. 1998, Icarus, 132, 125
Article number, page 9 of 9
|
Subsets and Splits